ELUCIDATION  OF  ILBP  FAMILY  FOLDING  PATHWAY  AND  STUDY  OF  REENGINEERING  THEM  AS   FLUORESCENT  PROTEIN  TAGS  VIA  STRUCTURAL  ANALYSIS     By    Zahra  Assar                       A  DISSERTATION     Submitted  to    Michigan  State  University     in  partial  fulfillment  of  the  requirements     for  the  degree  of       Chemistry  -­‐  Doctor  of  Philosophy       2018       ABSTRACT   ELUCIDATION  OF  ILBP  FAMILY  FOLDING  PATHWAY  AND  STUDY  OF  REENGINEERING  THEM  AS   FLUORESCENT  PROTEIN  TAGS  VIA  STRUCTURAL  ANALYSIS     By    Zahra  Assar   The  intracellular  lipid  binding  proteins  (iLBP)  family  are  found  in  the  cells  of  mammals,   birds,   fish,   amphibians   and   reptiles.   They   function   to   shuttle   large   insoluble   hydrophobic   molecules,  including  retinal,  various  long  chain  fatty  acids  and  etc.  throughout  the  cell  around   the   cytosol   and   nucleus.   The   combination   of   their   small   size   and   relatively   large   binding   pocket   make  them  suitable  templates  in  a  variety  of  protein  design  applications,  including  the  study  of   an  innovative  class  of  fluorescent  proteins.  To  pursue  our  goals,  we  used  human  Cellular  Retinol   Binding   Protein   II   (hCRBPII).   We   were   the   first   group   to   achieve   the   structure   of   an   all-­‐trans-­‐ retinal  and  the  first  bonafide  structure  of  retinol-­‐bound  hCRBPII.     In   the   course   of   these   studies,   we   have   discovered   hCRBPII   is   surprisingly   capable   of   folding  as  domain  swapped  dimer  (DS),  with  single  mutations  able  to  shift  the  folding  product   from  monomer  to  dimer.  Structural  analysis  of  both  wild  type  and    multiple  mutant  DS  dimers   led  us  to  remarkable  hypotheses  regarding  mechanism  of  this  phenomenon,  which  is  different   from   the   previous   studies   on   this   family.   We   proposed   that   the   N-­‐terminal   and   C-­‐terminal   halves   of   hCRBPII   are   capable   of   at   least   partially   folding   independently,   to   form   “open   monomer.”     The   dimer/monomer   ratio   depends   on   the   relative   rates   of   dimerization   of   the   open  monomers,  versus  closing  of  the  two  halves  together  to  form  the  “closed  monomer”.   In   addition,  by  comparing  structures  of  holo  hCRBPII  DS  dimer  variants,  we  identified  an  extremely   large  change  in  the  relative   orientation  of   the  two  domains   upon   ligand   binding  in   dimers.  This   suggests   the   possibility   that   iLBP   domain   swapped   dimers   could   be   allosterically   regulated   forms  of  these  proteins,  at  least  in  some  cases.  Fatty  acid  binding  protein  5  (FABP5),  another   member  of  iLBPs,  has  also  been  reported  to  forms  a  very  similar  DS  dimer,  which  makes  it  likely   that  other  family  members  could  also  form  DS  dimers,  and  have  physiological  importance  for   some  members  of  the  family.     As  mentioned,  a  new  class  of  fluorogenic  proteins  was  created  by  binding  fluorophore   aldehydes  in  the  binding  pocket  of  hCRBPII  via  protonated  Schiff  base  (PSB)  formation.  In  this   new   system,   emission   of   the   designed   solvatochromic   fluorophore   is   flexible   based   on   the   polarity  of  the  environment;  therefore  multicolor  probes  can  be  developed.  More  importantly,   absorption/emission  wavelengths  can  be  tuned  therefor;  nonspecific  labeling  and  background   fluorescence  can  be  reduced.  By  now,  the  absorption  maxima  are  tuned  from  501nm  to  705nm   and  emission  maxima  from  613  nm  to  744  nm.  Covering  both  the  red  and  far-­‐red  fluorescence   wavelength  regimes.                     Copyright  by     ZAHRA  ASSAR   2018                             This  Dissertation  is  lovingly  dedicated  to  my  grandmother,  Maman  joon.     v     ACKNOWLEDGMENTS     I  am  very  much  grateful  to  my  PhD  advisor,  Professor  James  H.  Geiger,  for  his  support   and  guidance  throughout  my  graduate  work.  He  taught  me  how  to  think  and  design  strategies   as  a  scientist.    He  also  trained  me  to  think  independently  and  to  be  more  patient  in  science   because  “Good  things  do  not  come  easy.  The  roak  is  lined  with  pitfalls”.  Desi  Arnaz.    Jim  always  supported  me  and  my  thoughts  about  my  projects.  Not  only  he  is  a  great  advisor,   also   he   is   very   friendly   approachable   person.   I   am   so   glad   and   honored   to   have   him   as   my   advisor.   Thank   you   Jim,   for   letting   me   to   grow,   even   start   my   professional   career   before   I   defend  and  thank  you  for  everything.     I   would   like   to   thank   my   second   reader   and   committee   member,   Professor   Babak   Borhan.  I  still  remember  August  2012  when  I  first  came  to  US,  I  was  not  sure  about  my  graduate   research   field   but   Babak   guided   me   very   well.     Also   during   biomeetings,   He   always   pointed   important   discussions   and   suggested   critical   experiments   which   they   all   helped   me   to   understand  my  research  better.  Whenever  I  needed  help  he  was  always  available  for  me.  Thank   you  Babak  for  everything.   I   would   like   to   appreciate   Dr.   William   Henry   for   being   on   my   committee   and   reading   my   dissertation.  I  received  some  great  advice  from  him  regarding  my  research.  In  addition,  thank   you  to  Professor  Gary  Blanchard  for  also  reading  my  dissertation.   I  would  like  to  thank  Dr.  Zahra  (Rafida)  Nossoni  and  Dr.  Wenjing  Wang  for  opening  a  path  for   me   to   pursue   my   PhD.   I   am   happy   that   I   had   Rafida   right   when   I   joined   the   Geiger   lab.   She   vi     helped  me  in  my  first  year  a  lot  and  she  was  always  available  especially  when  I  needed  help  in   data  collection.  She  is  a  great  sister-­‐in-­‐law  and  like  a  sister  that  I  have  never  had.   I   also   would   like   to   thank   Dr.   Chrysoula   Vasileiou   for   her   help.   She   always   was   available   and   helpful  to  me  with  her  ideas  and  great  discussions  in  our  biomeetings.    Thank  You  Chrysoula.   I   am   heartily   thankful   to   our   awesome   group   members,   all   current   and   former   members   in   Geiger’s  group  for  being  friendly  and  helpful  to  me.  Specifically,  thank  you  Dr.  Remie  Fawaz,  Dr.   Camille  Watson  and  Dr.  Meisam  Nosrati,  I  still  remember  my  first  Argonne  trip  and  how  much   Meisam   helped   me   during   my   first   time   data   collection.     Thank   you   Hadi   Nayebi   and   Alireza   Ghanbarpour   for   being   great   lab   members,   I   always   enjoyed   our   scientific   discussions   and   chats.   Also   thank   you   Nona   Ehyaei   for   always   cheering   me   up   with   your   energy   and   helping   me   whenever  I  could  not  drive  to  East  Lansing.  Thank  you  Lindsey  Gilbert  and  Janice  Chiou,  Kevin   Kramer   and   Colin   McCornack.     I   would   like   to   thank   the   bio   people   in   Professor   Borhan’s   group   Dr.  Tetyana  Berbasova,  Dr.  Ipek  Yapici,  Wei  Sheng  and  especially  Dr.  Elizabeth  Santos.  Liz  was   the  best  roommate  to  live  and  great  scientist  to  work  with.     I  would  like  to  thank  LS-­‐CAT  staff  at  the  Argonne  National  Laboratory  for  all  of  their  help  during   data   collection.   In   addition,   I   am   thankful   to   my   new   scientific   family   of   Cayman   Chemical   Company,  Dr.  Adam  Stein  and  Melissa  Holt.  I  am  very  happy  to  work  with  them  as  a  member  of   structural  biology  team.     I  also  appreciate  my  dear  friends  who  supported  me  on  this  path.     Finally,   I   am   grateful   to   my   beloved   family.   My   parents,   brother   and   my   husband   Dr.   Farid   Nossoni,   without   them,   their   love   and   support,   it   would   not   be   possible   for   me   to   finish   this   stage.   vii   Farid  always  taught  me  to  be  positive.  I  am  thankful  to  have  him  in  my  life,  for  inspiring  me  and   supporting  me  to  pursue  my  path.   And  again  I  would  like  to  say  thank  you  to  all  of  the  amazing  people  who  supported  and  helped   me.                                 ‘           viii   TABLE  OF  CONTENTS     LIST  OF  TABLES………………………………………………………………………………………………………………………...xii   LIST  OF  FIGURES……………………………………………………………………………………………………………………...xiii   KEY  TO  SYMBOLS  AND  ABBREVIATIONS………………………………………………………………………………....xxii   Chapter  I:  Structural  Studies  of  Human  Cellular  Retinol  Binding  Protein  II  (hCRBPII)  bound  to   Retinol  and  Retinal.  ................................................................................................................  1   I-­‐1  Introduction  ..........................................................................................................................  1   I-­‐2  The  affinity  of  hCRBPII  for  retinol  and  retinal  .......................................................................  4   I-­‐3  Structure  of  the  retinal-­‐bound  hCRBPII  complex  ..................................................................  5   I-­‐4  Wavelength-­‐dependent  damage  of  the  retinol  in  hCRBPII  ...................................................  8   I-­‐5  The  structural  difference  between  retinol-­‐bound  hCRBPII  and  retinal-­‐bound.  ..................  13   I-­‐6  Ligand  Binding  in  CRBPI  versus  CRBPII  ................................................................................  15   I-­‐7  Conclusion  ...........................................................................................................................  17   I-­‐8  Experimental  .......................................................................................................................  18                      I-­‐8-­‐1  Material  and  Method:  Protein  Expression  and  Purification  ...................................  18                      I-­‐8-­‐2    Kd  Determination  via  Fluorescence  Quenching  Assay  ...........................................  19                      I-­‐8-­‐3  Crystallization  and  Structure  Determination  ..........................................................  20   REFERENCES  .............................................................................................................................  24   Chapter  II:  Domain  swapping  in  hCRBPII  ................................................................................  28   II-­‐1  Introduction  .......................................................................................................................  28   II-­‐2  Origin  of  dimerization  in  hCRBPII  .......................................................................................  35   II-­‐3  Dimer  formation  of  other  hCRBPII  variants  .......................................................................  39   II-­‐4  Structural  Studies  ...............................................................................................................  44                      II-­‐4-­‐1  Structural  analysis  reveals  an  extensive  domain  swapped  dimer  .........................  44                      II-­‐4-­‐2  Symmetry  VS  asymmetry  .......................................................................................  48                      II-­‐4-­‐3  Phase  Relationship  .................................................................................................  53   II-­‐5  possible  mechanism  for  domain  swapping  in  hCRBPII  .......................................................  55                      II-­‐5-­‐1  Dimerization  and  structure  of  the  E72A  mutant  ...................................................  58                      II-­‐5-­‐2  Study  the  folding  pathway  .....................................................................................  60                      II-­‐5-­‐3  HCRBPII  folding  route  VS  other  iLBP  members.  .....................................................  63   II-­‐6  Conformational  change  driven  by  ligand  binding  in  hCRBPII  .............................................  64   II-­‐7  Experimental  ......................................................................................................................  73                      II-­‐7-­‐1  Material  and  Method:  Site-­‐Directed  Mutagenesis  ................................................  73                      II-­‐7-­‐2  Protein  Expression  and  Purification  of  DS  dimer  mutants  .....................................  76                      II-­‐7-­‐3  Protein  Refolding  ...................................................................................................  77   ix                        II-­‐7-­‐4  Thermal  Melting  Curves  for  monomer  and  dimer  mutants.  ..................................  78                      II-­‐7-­‐5    Kd  determination  ...................................................................................................  80                      II-­‐7-­‐6  Extinction  Coefficient  Determination  ....................................................................  80                      II-­‐7-­‐7  Crystallization,  data  collection  and  refinement  .....................................................  80   REFERENCES  .............................................................................................................................  85   Chapter  III:  Can  we  predict  and  control  the  folding  product  of  proteins  through  their  amino   acid  sequence    (at  least  in  iLBP  family)?  .................................................................................  95   III-­‐1  Introduction  ......................................................................................................................  95   III-­‐2  Investigation  of  potential  domain  swapping  in  other  iLBP  family  members:  Case  Study  1,   human  FABP4  ...........................................................................................................................  96                      III-­‐2-­‐1  History  of  FABP4  ...................................................................................................  96                        III-­‐2-­‐2  Investigation  on  hFABP4  bacterial  expression  ......................................................  97   III-­‐3  Investigation  of  potential  domain  swapping  in  other  iLBP  family  members:  Case  Study  1,   human  FABP5  ...........................................................................................................................  98                      III-­‐3-­‐1  History  of  hFABP5  .................................................................................................  98                      III-­‐3-­‐2  Investigate  domain  swapping  in  human  fatty  acid  binding  protein  5  .................  103                      III-­‐3-­‐3  Study    domain  swapping  occurance  by  mutational  analysis  of  human  FABP5  ...  108   III-­‐4  Experimental  ...................................................................................................................  111                      III-­‐4-­‐1  Material  and  Method:  Site-­‐Directed  Mutagenesis  .............................................  111                      III-­‐4-­‐2  Protein  Expression  and  Purification  of  hFABP4  and  hFABP5…………………………….113                      III-­‐4-­‐3  Crystallization,  data  collection  and  refinement  ..................................................  114          REFERENCES  ...........................................................................................................................  116     Chapter  IV:  STRUCTURAL  INSIGHTS  INTO  EMISSION  REGULATION  OF  HCRBBPII-­‐BOUND   SOLVATOCHROMIC  FLUOROPHORE  .....................................................................................  127   IV-­‐1  Introduction  ....................................................................................................................  127   IV-­‐2  Rational  behind  the  parent  template:  Q108K:K40L:T51V:T53S  ......................................  128   IV-­‐3  Attempts  to  isolate  the  binding  cavity  ............................................................................  132   IV-­‐4  Extensive  packing  of  the  ligand  and  removal  of  water  leads  to  bathochromic  shift  in   absorption  and  emission.  .......................................................................................................  136   IV-­‐5  Localization  of  charge  leads  to  a  blue  shift  in  absorption  and  emission  ........................  138   IV-­‐6  Material  and  Methods  ....................................................................................................  141   REFERENCES  ...........................................................................................................................  144     Chapter  V:  Mutational  Studies  on  Starch  Branching  Enzyme  I  (RBEI)  to  understand  its  Surface   Binding  Mechanism.  ............................................................................................................  157   V-­‐1  introduction  .....................................................................................................................  157   V-­‐2  Structure  of  the  both  N  and  C  terminus  truncated  RBEI:  ................................................  159                      V-­‐2-­‐1  Binding  Site  I:  Carbohydrate  Binding  Module  (CBM)  ...........................................  162                      V-­‐2-­‐2  New  Observed  Binding  Site  II  ..............................................................................  163   V-­‐3  Mutational  studies  on  the  observed  binding  sites  and  active  site  of  RBEI  ......................  164   x                        V-­‐3-­‐1  Effect  of  Active  Site  Mutation:  ............................................................................  165                      V-­‐3-­‐1  Effect  of  E534A  Mutation:  ...................................................................................  167   V-­‐4  Experimental  ....................................................................................................................  170                      V-­‐4-­‐1  Material  and  Method  ..........................................................................................  170                      V-­‐4-­‐2  Activity  Assay  .......................................................................................................  173                      V-­‐4-­‐3  Crystallization  of  RBEI  mutants  ............................................................................  175   REFERENCES  ...........................................................................................................................  176                                           xi   LIST  OF  TABLES     Table  I-­‐1:      Data-­‐collection  and  refinement  statistics...................................................................22   Table  II-­‐1:    The  Effect  of  expression  temperature  and  in  vitro  refolding  protein  concentration  on                                            the  dimer/monomer    (D/M)  ratio...............................................................................40     Table  II-­‐2:  Monomer/dimer  ratio  of  various  residue  60  mutants.  During  bacterial  expression  at                30  °C............................................................................................................................42     Table  II-­‐3:      Summary  of  DS  dimer  crystal  structures  of  all  hCRBPII  mutants...............................66     Table  II-­‐4  :    PCR  protocol  for  hCRABPII  and  hCRBPII  mutagenesis...............................................74     Table  II-­‐5:    Anion  Exchange  purification  protocol  for  hCRBPII.  The  buffer  used  for  above                                            protocol  is  50  mM  Tris,  pH  is  adjusted  automatically.  The  proteins  elute  between  4%                                          -­‐8  %  2M  NaCl.  .............................................................................................................77     Table  II-­‐6:  Calculated  percentage  of  helix/strand  for  each  denaturant  (GD.HCl)  concentration                                          during  unfolding  experiment.  220nm  corresponds  to  beta  sheets.  222nm/208nm                                          corresponds  to  ι-­‐helixes.............................................................................................80     Table  II-­‐7:    X-­‐ray  crystallographic  data  and  refinement  statistics  for  hCRBPII    mutants..............82     Table  III-­‐1:    Crystallographic  data  of  hFABP5  (pseudo  monomer).............................................115     Table  IV-­‐1:  Spectroscopic  change  of  hCRBPII  variants  bound  ThioFluor....................................132     Table  IV-­‐2:    X-­‐ray  crystallographic  data  and  refinement  statistics  for  monomeric  hCRBPII                                              mutants....................................................................................................................143     Table  V-­‐1:    Residues  involved  in  the  interactions  in  Binding  sites  I,  II  and  III  of  RBEI-­‐                                            M12........................................................................................................................165                                 Table  V-­‐2:    More  suggested  mutations  to  study  for  residues  interacting  with  M12  in  binding                                          site  II  of  RBEI.  ............................................................................................................170         xii   LIST  OF  FIGURES       Figure  I-­‐1:    a.  An  overlay  the  structures  of  hCRBPII  (green  and  cyan),  hCRBPIII  (1GGL,  salmon),                  hCRBPIV  (1LPJ,  yellow)  and  hCRBPI  (1KGL,  pink)  b.  An  overlay  of  the  structures  of  apo                          hCRBPII  (2RCQ,  gray),  wt  hCRBPII  bound  to  all-­‐trans-­‐retinal  (4QYP,  green)  and  to  all-­‐                      trans-­‐retinol  (4QYN,cyan)………………………………………………………………………………………….2     Figure  I-­‐2:  The  multiple  sequence  alignment  between  human,  rat  CRBPs  and  zebra  fish  CRBP...3     Figure  I-­‐3:  Ligand  binding  measurements  using  the  fluorescence  quenching  of  tryptophan                           (excitation   at   280   nm,   emission   at   350   nm).   a.  Titration   of   WT-­‐hCRBPII   with   all-­‐trans-­‐         retinol.   b.   Titration   of   hCRBPII   T51I   mutant   with   all-­‐trans-­‐retinol.   c.   Titration   of   hCRBPII  T51I  mutant  with  all-­‐trans-­‐retinal.    d.  Titration  of  WT-­‐hCRBPII  with  all-­‐trans-­‐ retinal.  ………………………………………………………………………………………………………………………5     Figure  I-­‐4:    Retinal  binding  in  hCRBPII.  a.  An  overlay  of  retinal  from  mol  A  (cyan),  mol  B  (blue)         and  Mol  C  (pink),  showing  the  similarity  in  binding  in  the  three.  b.  Stick         representation  of  the  interaction  between    Phe16,  Leu77  (both  with  green  carbons,  N                      blue  and  O,  red)  and  the  retinal  ionone  ring  (blue  carbons,  Mol  B).  c.  Space  filling                depiction  of  b.  ………………………………………………………………………………………………………..…7     Figure  I-­‐5:    The  water  mediated  interaction  of  the  retinal  hydroxyl  group  in  hCRBPII.    Atoms                       colored  by  type,  hydrogen  bonded  distances  are  shown.………………………………………….8     Figure  I-­‐6:  a.  The  contoured  2Fo-­‐Fc  electron  density  map  at  1.0  σ  for  rearranged  retinol  from                                            data  collected  at  11  KeV/the  highest  flux  (4QYN,  blue).  b.  The  chemical  structure  of  a                                        possible  retinoid  derivative  consistent  with  the  crystallographic  data.  c.  The  overlay  of                                            bonafide  retinol  structure  (chain  C)  from  data  collected  with  an  attenuated  X-­‐ray              beam  at  an  energy  of  7  KeV  (4QZT,  Pink)  and  the  structure  of  rearraged  retinoid                                        derivative  (chain  B)  from  data  collected  at  11  KeV  and  the  highest  flux  (4QYN,  blue).                                        d.  The  structures  of  the  noncanonical  retinoid  (refined  in  two  conformations)                                        previously  seen  in  the  human  CRBPII  (2RCT,  yellow),  and  our  current  rearranged                                        structure  (4QYN,  blue)  were  overlaid.  e.  The  structures  of  the  canonical  and                                        noncanonical  retinoid  previously  seen  in  the  zebrafish  CRBP  (1KQW,  green),  and  our                                        current  rearranged  structure  (4QYN,  blue)  were  overlaid.  f.  The  overlaid  structures  of                                        the  canonical  retinol  obtained  in  the  rat  CRBPII  (1OPB,  purple)  and  our  current                                        bonafide  all-­‐trans-­‐retinol  structure  in  human  CRBPII  (4QZT,  pink).…………………………….9     Figure  I-­‐7:    a.  The  rearranged  retinol  from  data  collected  at  11  KeV  and  the  highest  flux  (4QYN,                blue)  and  the  interaction  of  alcohol  moiety  with  the  neighboring  residues  of  Gln108                                          and  Lys40.  b.  The  structures  of  the  retinal  bound  hCRBPII  (4QYP,  cyan,  chain  B),  and                                               xiii   our   current   rearranged   retinol   structure   (4QYN,   blue)   were   overlaid.   The   order   water   molecule   (W)   seen   in   all-­‐trans-­‐retinal   crystal   structure,   is   occupied   by   the   hydroxyl   group   of   the   noncanonical   retinoid   (indicated   in   dashed   circle).   Similar   case   in   retinol   structure   from   data   collected   at   7   KeV   and   an   attenuated   beam   has   been   seen.   c.   Interactions  around  Gln108  of  retinol  bound  rat  CRBPII  (1OPB,  purple)..………………...10     Figure  I-­‐8:    This  experiment  performed  by  Dr.  Yapici  a.  UV-­‐vis  spectrum  of  all-­‐trans-­‐retinol.    b.                                              The  HPLC  trace  of  all-­‐trans-­‐retinol.  c.    UV-­‐vis  spectrum  of  the  retinol  extracted  from                                              the  crystals  and  purified  by  HPLC.  d.  HPLC  chromatogram  of  the  retinol  extracted                                            from  crystals.  e.    HPLC  chromatogram  of  bonafide  all-­‐trans-­‐retinol  and  the  retinol                                            extracted  from  crystals.  Showing  that  they  are  the  same  species.............................12     Figure  I-­‐9:  A  comparison  of  the  retinol  and  retinoid  moieties  found  in  the  crystal  structures  of                                          human,  zebra  fish  and  rat  CRBPII.  a.  The  retinol  (left)  and  retinoid  (right)  moieties                                              found  in  the  previously  published  hCRBPII  (2RCT,  yellow).  The  value  of  128°  torsion                                          angle  (ψ)  about  the  C13-­‐C14  bond  is  inconsistent  with  a  double  bond  (180°).  b.    The                                          retinol  and  retinoid  moieties  seen  in  the  crystal  structure  of  zebrafish  CRBP  (1KQW,                                          green).    The  torsion  angle  about  the  C13-­‐C14  bond  is  also  inconsistent  with  a  double                                          bond.  c.  The  retinol  structure  in  the  rat  CRBPII  (1OPB,  purple).  d.    The  retinol  found  in                                          the  X-­‐ray  structure  of  hCRBPII  using  data  collected  at  an  energy  of  7  KeV  and  an                                          attenuated  beam  (4QZT,  pink)  with  an  approximate  180°  torsion  angle  around  C13-­‐                                        C14  bond.  …………………………….......................................................................................14     Figure  I-­‐10:  Retinal  versus  retinol  binding  in  hCRBPII.  a.  The  structure  of  the  retinol-­‐bound  WT                                              hCRBPII.  b.  The  2Fo-­‐Fc  electron  density  map  contoured  at  1.0σ  around  retinol  of                                              7Kev  data.  c.  The  structure  of  all-­‐trans-­‐retinal  bound  to  WT  hCRBPII.  d.  The                                              structures  of  all-­‐trans-­‐retinal  (green)  and  retinol  (magenta).  The  position  of  an                                              ordered  water  molecule  (W)  occupied  by  the  hydroxyl  group  of  retinol  is  indicated                                                in  dashed  circle.........................................................................................................15     Figure  I-­‐11:  The  structures  of  retinol-­‐bound  rat  CRBPI  (magenta  carbons  and  hydrogen  bonds)                                                and  retinal-­‐bound  hCRBPII  (green  carbons  and  hydrogen-­‐bonds)  are  overlayed.                                                Amino  acid  labels  are  colored  similarly  when  different  in  the  two  structures.    Note                                                the  change  in  the  water  network  introduced  by  Phe3…...........................................16     Figure  II-­‐1:  Cartoon  illustration  of  3D  domain  swapping.  Identical  structural  units  swap  between                                            2  monomers,  followed  by  creating  the  open  interface  between  them......................28     Figure  II-­‐2:  a.  Left:  DT  monomer  (PDB  ID  4OW6).  Right:  DT  dimer  (4AE1).  The  two  subunits  are                                          blue  and  pink.  In  acidic  pH  equilibrium  goes  toward  DS  dimer  formation.  b.  Left                                          structure  of  DS  dimer  of  RnaseA  (PDB  1A2W)  and  cyclic  DS  trimes  (PDB  5SRA).........30     Figure  II-­‐3:    The  hCRBPII  DS  dimer................................................................................................34     xiv   Figure  II-­‐4:  Trajectory  of  Tyr  60,  Glu  72  (both  shown  as  stick  representations,  colored  blue)  in                                                WT  hCRBPII  with  all-­‐trans-­‐retinal...............................................................................36     Figure  II-­‐5:  a.  IEX  chromatogram  of  the  KLY60W,  monitored  at  280nm.  b.  The  UV-­‐Vis  spectra  of                                                                      the  80mM  salt  elution  (red)  and  the  160mM  salt  elution  (blue)  of  KLY60W  incubated                                              with  all-­‐trans-­‐retinal.  c.  The  UV-­‐Vis  spectra  of  the  refolded  80mM  eluent  (low  salt,                                            blue)  after  denaturation  and  refolding  at  0.03  mM  concentration  and  the  native                                            160mM  (high  salt,  green)  and  native  80mM  elution  (red),  all  bound  to  all-­‐trans-­‐                                          retinal.  d.  The  size  exclusion  chromatogram  (Superdex  200  16/75  column)  for  dimer                                            (the  160mM  elution)  KLY60W,  after  maintaining  the  protein  at  room  temperature                                            for  several  days,  showing  only  the  monomer  size  peak.  e.  Gel  filtration                                            chromatogram  of  monomer  (the  80mM  elution)  of  KLY60W,  showing  only  the    dimer                                              fractions.  f-­‐i.  Ion  exchange  chromatographs  of  WT  and  various  hCRBPII  mutants,                                            monitored  at  280  nm.  f.  Y60I  at  30°C.  g.  Y60L  at  30°C.  h.  Y60L  at  25°C.  i.  WT  hCRBPII                                            expressed  at  30  °C......................................................................................................37     Figure  II-­‐6:    Thermal  melting  curves:  a.  WT  hCRBPII  dimer  (θ  measured  at  220  nm),  Tm  =  52  °C;                                              b.  Y60L    dimer  (θ  measured  at  227  nm),  Tm  =  56  °C;  c.  KLY60W  dimer  (θ  measured  at                                              220  nm),  Tm  =  69°C.  d.  WT  hCRBPII  monomer  (θ  measured  at  220  nm);  e.  Y60L                                              monomer  (θ  measured  at  215nm);  f.  KLY60W  monomer  (θ  measured  at  220  nm).                                              Note  that  most  of  the  secondary  structure  is  preserved  in  both  the  WT  and  KLY60W                                              monomers,  even  at  the  highest  temperature  achieved  (CD  intensities  of-­‐65mdeg                                              and  -­‐50mdeg  respectively).  Thus  melting  curves  can  only  give  a  lower  limit  for  the                                              TM,  consistent  with  the  view  that  the  monomers  are  more  stable  than  the  dimers.                                              The  apparent  lack  of  a  Tm  for  the  monomers  might  suggest  the  presence  of                                              thermally  induced  folding  intermediates.  Clearly  the  unfolding  process  is  not                                              monolithic  as  that  observed  for  the  dimeric  species.  This  lends  further  support  to                                              the  suggestion  that  the  open  monomer,  which  is  presumably  obtained  from  melting                                              of  the  dimer,  has  a  different  folding  trajectory  (and  consequently  a  different                                              unfolding  trajectory)  as  compared  to  the  monomeric  species..................................39     Figure  II-­‐7:    Size  exclusion  chromatography  after  protein  denaturation  and  refolding,  all                                              monitored  at  280nm.      a.  Y60L  mutant  refolded  at  0.03  mM  concentration.  b.  Y60L                                              refolded  at  0.13  mM  concentration.  c.  Y60L  refolded  at  0.4  mM  concentration.  d.                                                WT  hCRBPII  refolded  at  0.03  mM  concentration.  e.  WT  hCRBPII  refolded  at  0.13  mM                                              concentration.  f.  WT  hCRBPII  refolded  at  0.3  mM  concentration.  ...........................41     Figure  II-­‐8:    a.The  size  exclusion  chromatogram  for  dimer  of  WT  hCRBPII,  after  maintaining  the                                              protein  at  room  temperature  for  several  days,  showing  only  the  dimer  species.  b.                                                Gel  filtration  chromatogram  of  monomer  of  WT,  showing  only  the  monomer                                              protein.......................................................................................................................43     Figure  II-­‐9:    Ligand  binding,  monitored  by  Tryptophan  fluorescence  quenching  of  monomeric                                              and  dimeric  hCRBPII.  a.  Monomer  WT  hCRBPII  with  all-­‐trans-­‐retinol.  b.  Dimer  WT     xv                                            hCRBPII  with  all-­‐trans-­‐retinol.  c.  Monomer  WT  hCRBPII  with  all-­‐trans-­‐retinal.  d.                                              Dimer  WT  hCRBPII  with  all-­‐trans-­‐retinal....................................................................44     Figure  II-­‐10:  Crystal  structure  of  dimer  KLY60W  mutant  and  its  monmer  components  a.  Chain  A                                                  of  KLY60W  (where  only  the  Thr56  psi  angle  differs  from  the  closed  monomer  form)                                                  with  the  electron  density,  contoured  at  1.0  σ,  of  the  hinge  loop  region                                                      encompassing  amino  acids  55-­‐60.    Atoms  colored  by  type,  N,  blue,  O,  red,  C,  cyan.                                                  b.  Chain  B,  showing  the  electron  density,  contoured  at  1.0  σ,  of  the  hinge  loop                                                  region  encompassing  amino  acids  55-­‐60.  Atoms  colored            by    type,  N,  blue,  O,  red,                                                  C,  cyan.  c.  The  complete  structure  of  the  KLY60W  dimer:  In  chain  A    (cyan)  Trp60  is                                                  pointing  inside  the  binding  pocket,  while  in  chain  B  (pink)  Trp60  is  pointing  toward                                                  the  solvent...............................................................................................................45     Figure  II-­‐11:    An  overlay  of  the  dimeric  (cyan)  and  monomeric  (red)  KLY60W  mutant  structures.                                                  Res56  for  both  structures  is  shown  in  sticks.  The  hinge  loop  region  in  the  monomer                                                  and  dimer  is  encircled,  showing  the  dramatic  difference  in  the  two  structures......46     Figure  II-­‐12:    a-­‐e:  Torsion  angle  differences  between  monomer  and  dimer  differences  of  Phi                        angle  are  in  blue  and  psi  angle  differences  are  in  red  in  each  case.    a.  Thr56.  b.                        Phe57.  c.  Arg58.  d.  Asn59.  e.  Res60………………………………………………………………..……47     Figure  II-­‐13:  The  differences  in  the  asymmetric  and  symmetric  structures  of  hCRBPII  dimers                                                  Res60  in  all  structures  is  shown  in  sticks  a.  The  asymmetric  structures:  The  overlay                                                  structure  of  domain  swapped  dimer  KLY60W  (cyan)  and  Y60W  (pink).  In  chain  A                                                  dimer  KLY60W,  Trp60  is  pointing  inside  the  binding  pocket.  b.  The  symmetric                                                  structures.  The  overlay  of  WT  hCRBPII  (purple)  and  Y60L  (yellow)  domain  swapped                                                  dimer.  c.  Comparing  asymmetric  (dimer  KLY60W  in  cyan)  and  symmetric  dimer                                                  (dimer  Y60L  in  yellow)  by  looking  down  the  two-­‐fold  axis  and  it  shows  the  different                                                  position  of  helices  in  N-­‐terminus.  d.  One  chain  of  dimer  Y60L  (yellow)  and  both                                                  chains  of  asymmetric  dimer  Y60W  (chain  A  in  cyan  and  chain  B  in  red)  are  overlaid                                                  at  their  C-­‐terminus...................................................................................................48     Figure  II-­‐14:    Torsion  angle  deviations  outside  the  hinge-­‐loop  region  define  the  relative                                                          orientation  of  the  two  domains  of  the  dimer,  with  the  highest  deviation  seen  in                                                  Asn59.  a.    The  WT  (purple)  and  Y60L  (yellow)  dimers  are  overlaid.    Inset,  the                                                  hydrogen  bond  between  Tyr60  and  Glu72  compared  to  the  “flipped  out”                                                  conformation  of  Leu60.  b.  The  critical  Tyr60  and  Asn59  region  in  the  WT  hCRBPII                                                  dimer,  showing  the  key  phi/psi  angles.  c.  The  same  region  in  the  Y60L  dimer,                                                  showing  the  “flipped  out”  Leu60  and  the  key  phi/psi  angles.    Comparison  of  the                                                  two  shows  how  the  large  difference  in  the  phi  angle  is  compensated  for  in  the  WT                                                  N59  psi  angle,  keeping  the  main  chain  of  the  two  on  a  similar  trajectory……………50      Figure  II-­‐15:  Residues  56-­‐63  in  the  chains  A  (green,  top)  and  B  (magenta,  bottom)  of  Y60W                                                  hCRBPII.....................................................................................................................51   xvi     Figure  II-­‐16:  Overview  of  dimer  formation  in  asymmetric  structures:  Two  chains  of  Y60W  dimer                                              are  labeled  in  the  picture.  The  same  chain  (molecule  A  and/or  B)  in  this  asymmetric                                                dimers  cannot  form  dimers  since  they  clash.  However,  the  interaction  of  different                                              chain  with  each  other  (molecule  A  and  B)  could  lead  to  proper  DS  dimer  formation                                              with  the  different  orientation....................................................................................53     Figure  II-­‐17:  Schematic  representation  of  “phase  relationship”  in  DS  dimerization.  Top:                                                monomer,  bottom:  dimer  with  residue  60  as  mutant..............................................54     Figure  II-­‐18:    DS  dimerization  requires:  1.  Rotation  about  Thr56  psi.  2.  Orientation  of  the  two                                                  halves  to  accommodate  dimerization.    3.  Rephasing  of  the  connecting  strand.......56     Figure  II-­‐19:    DS  dimer  vs.  monomer  formation  in  WT  hCRBP  II..................................................57     Figure  II-­‐20:    Size  exclusion  chromatography  of  low  salt  and  high  salt  (80  mM  and  160  mM                                                  NaCl,)  E72A  hCRBPII  respectively,  expressed  at  25°C  (since  30°C  expressed                                                  produced  virtually  no  soluble  protein).....................................................................58     Figure  II-­‐21:  a.  Overlay  of  chain  A  of  the  E72A  dimer  (green)  and  the  Y60L  dimer  (yellow).  Note                                                that  both  show  residue  60  pointing  toward  the  solvent.  b.  Overlay  of  the  E72A                                                chain  A  (blue)  and  WT  (green)  dimers.  Inset  shows  that  without  the  Glu  72                                                hydrogen  bond  donor,  residue  60  is  free  to  flip  out,  leading  to  a  more  relaxed                                                dimer  structure.  c.  Two  chains  of  the  E72A  dimer  (green  and  grey)  are  overlaid,                                                showing  their  strong  similarity.................................................................................60     Figure  II-­‐22:  Equilibrium  Unfolding  Experiment  for  wild-­‐type  CRBPII  in  presence  of  different                                                concentration  of  Gd-­‐HCl,  a.-­‐b.  monitoring  by  Circular  Dichroism  (CD).  c-­‐d.                                                Tryptophan  fluorescence  spectroscopy,  right,  performed  with  WT  hCRBPII,  to  study                                                the  presence  an  intermediate  in  the  folding  pathway.    Purple  solid  line  represents                                                the  polynomial  fit  (3rd  degree)  in  b  and  d.  Both  b  and  d  spectra  are  in  correlate  with                                                previous  studies.  (1)    .................................................................................................62     Figure  II-­‐23:    overview  picture  of  hCRABPI  folding  pathway  compare  to  hCRBPII.......................64     Figure  II-­‐24:    A  comparison  of  four  symmetric  dimers,  Y60L  (orange),  WT  (red),  E72A  (cyan)  and                                                                                                    Q108K,K40D  (green)...............................................................................................,..66     Figure  II-­‐25:  Overlay  of  holo  Q108K:K40D  (chains  F  and  A,  both  shown    in  orange)  and  apo                                                Q108K:T51D  (shown  in  blue)  showing  the  large  motion  of  helix  1  upon  ligand                                                binding  (about  15  Å).  Bound  retinal  molecules  in  holo  structures  are  shown  as                                                transparent  spherical  models.  Note,  complete  dimer  in  the  apo  structures  was                                                generated  by  crystallographic  two-­‐fold  symmetry  operation.  ................................67     xvii   Figure  II-­‐26:    Ligand  binding  leads  Asn59  to  flip  out  of  the  binding  pocket  due  to  steric  effect.  a.                                                  Overlay  of  Q108K:T51D-­‐apo  (green,  obtained  by  ALireza)  and  Q108K:K40D-­‐holo                                                  (orange).  Allosteric  conformational  change  between  apo  and  holo  dimers  is  caused                                                  by  the  orientation  of  the  Tyr  60  and  Asn  59  sidechains.  Asn59  would  sterically  clash                                                  with  the  bound  retinylidene,  which  leads  to  the  flipped-­‐out  conformation  of  Asn59                                                  and  flipped-­‐in  conformation  of  Tyr60.  b.  Overlay  of  WT  apo  monomer  (pink,  PDB                                                  code:  2RCQ)  and  WT  apo  hCRBPII  dimer  (green,  PDB  code:  4ZH9).  The  “flipped  in”                                                  Asn59  conformation  is  required  to  adopt  the  orientation  to  form  the  symmetric  DS                                                    dimer.  c.  The  similar  trajectory  of  retinal  in  monomer  and  dimer  structures  of                                                    hCRBPII  mutants.  Q108K:K40L  monomer  (purple,  PDB  code:  4RUU)  Q108K:K40D                                                    dimer  (orange)  and  Q108K:K40L:T51DD  dimer  (pink,  obtained  by  Dr.Nosrati)  which                                                    are  all  bound  to  retinal  via  a  covalent  bond  to  Lys108.  d.  Overlay  of  WT  monomer                                                    hCRBPII  bound  to  retinol  (green,  PDB  code:  4QZT)  with  retinal-­‐bound  Q108K:K40D                                                    dimer  (orange)........................................................................................................68     Figure  II-­‐27:  Overlay  of  WT-­‐hCRBPII  (holo)  DS  dimer  shown  by  dark  blue  and  WT-­‐hCRBPII  (apo)                                                  (PDB  code:4ZH9)  shown  by  green............................................................................69     Figure  II-­‐28:  Overlay  of  different  chains  in  holo  Q108K:K40D  DS  dimer.  a.  Chains  A  shown  by                                                  orange  and  I  shown  by  green.  b.  Chains  A  shown  by  orange  and  L  shown  by  blue.                                                    c.  ChainA  shown  by  orange  and  D  shown  by  pale  purple  d.  Chains  A  shown  by                                                  orange    and  G  shown  by  purple.  e.  Overlay  of  chains  A  shown  by  orange  and  B                                                      shown  by  yellow.  f.  Overlay  of  chains  A  shown  by  green  and  G  shown  by  cyan.  g-­‐l:                                                  E72…Y60  Hydrogen  boning  motifs  in  Q108K:K40D  holo  DS  dimer:  g.  Chains  A  and  F                                                        h.  Chains  D  and  H.    i.  Chain  L  and  J.  j.  Chains  B  and  E  k.  Chains  K  and  I.  l.  Chains  C                                                    and  G.......................................................................................................................71     Figure  II-­‐29:    UV  data  spectrum  for  different  dimer  mutants  with  all-­‐trans  retinal  was  carried  by                                                  Dr.  Santos.  Left,  UV  spectra  represents  the  kinetically  formed  product  between  the                                                  hCRBPII  dimer  and  retinal.  Right,  this  panel  depicts  the  formation  of  both  SB  and                                                  PSB  as  a  function  of  time.  ........................................................................................73     Figure  II-­‐31:    Circular  Dichroism  spectra  of  various  hCRBPII  mutants  as  a  function  of                                                    temperature.  a.  Monomer  KLY60W.  b.  Dimer  KLY60W.  c.  Monomer  Y60L.  d.  Dimer                                                    Y60L.  e.  Monomer  WT  hCRBPII.  f.  Dimer  WT  hCRBPII.    Note,  most  of  the  secondary                                                    structure  is  preserved  in  both  the  monomers,  even  at  the  highest  temperature                                                    achieved.    Thus  melting  curves  can  only  give  a  lower  limit  for  the  TM,  consistent                                                    with  the  view  that  the  monomers  are  more  stable  than  the  dimers......................79     Figure  III-­‐1:  Overlaid  of  FABP4  (2q9s,  green)  -­‐Linoleic  acid  (blue),  FABP4  (2ans,  pink)-­‐oleic  acid                                              (grey)  shows  the  exit  of  the  oleic  acid  from  the  binding  pocket...............................96     Figure  III-­‐2:  Size  Exclusion  Chromatogram  of  human  FABP4,  which  was  expressed  in  E.coli  (at                                              30°C  as  dimer  favoring  temperature,  explained  in  section  III-­‐3,  green).  SEC  profile  of     xviii                                            monomer  hCRBPII  mutant  (orange)  as  our  size  control............................................98     Figure  III-­‐3:    Model  for  the  role  of-­‐FABP5  and  its  involvement  in  regulating..............................99     Figure  III-­‐4    Overlaid  of  DS  dimers  of  holo  hFABP5  and  hCRBPII.  HCRBPII  (purple,  PDB  ID  4ZH9),                                                  BMS-­‐hFABP5  (green),  AEA-­‐hFABP5  (orange).  There  is  a  big  different  orientation  of                                                two  domains  upon  ligand  binding.  .........................................................................100     Figure  III-­‐5:  a.  Overlaid  of  holo  DS  dimers  of  BMS-­‐hFABP5  (green),  AEA-­‐hFABP5  (orange).                                              Dramatic  change  is  seen  in  the  relative  orientation  of  two  domains  in  these  DS                                              dimers,  which  is  induced  by  different  ligands.  b.  Overlaid  of  DS  dimers  of  BMS-­‐                                            hFABP5  (green),  AEA-­‐hFABP5  (orange)  and  holo  Q108K:K40D  DS  dimer  of  hCRBPII.                                              AEA,  BMS  and  retinal  are  shown  as  sticks.  Between  structures  of  dimer  hCRBPII  and                                              AEA-­‐hFABP5,  there  is  14Å  deviation  while  for  BMS-­‐hFABP5  dimer  there  is  a  larger                                              deviation  of  17Å.......................................................................................................102     Figure  III-­‐6:  a.  SEC  of  WT  and  Q4A  mutant  FABP5  purification  by  SEC  after  E.  coli  expression.    b.                                              SDS-­‐PAGE  of  SEC  fractions  18-­‐21  of  WT-­‐hFABP5.  c.    Native  gel  of  the  same  SEC                                              fractions.  It  shows  that  in  the  dimer  peak,  both  monomer  and  dimer  species  are                                              present.....................................................................................................................104     Figure  III-­‐7:  Overlaid  of  our  current  monomer  structure  (chain  A  in  green  and  chain  B  in  pink)                                              with  hFABP5  monomer  (grey,  PDB  ID  4lkp).  There  is  a  different  conformation                                              between  their  3rd  and  4th  beta  sheet  (hinge  loop  region  in  DS  dimer),  which  is                                              pointed  by  arrow.  ...................................................................................................105     Figure  III-­‐8:  a.  Crystal  structure  of  DS  dimer  of  hFABP5  of  our  trial.  Chain  A  in  blue,  Chain  B  in                                              cyan.  b.  Overlaid  of  DS  hFABP5  structure    (obtained  by  Nona,  cyan)  with                                              abovementioned  pseudo  monomer  hFABP5  (chain  A  in  pink  and  chain  B  in  green)c.                                              Overlay  of  our  determined  DS  hFABP5  structure    (blue)  with  DS  dimer  of  BMS-­‐                                            hFABP5  (green,  4AZM),  and  AEA-­‐hFABP5  (orange,  4AZR).    This  dimer  structure  is                                              very  similar  to  the  AEA-­‐  structure  and  both  have  different  orientation  of  their                                              domains  compared  with  the  BMS-­‐structure............................................................107     Figure  III-­‐9:  a.  In  DS  dimer  of  hFABP5  (Chain  A  in  salmon,  chain  B  in  cyan,  PDB  4AZR),  there  is  a                                              hydrogen  bond  between  Q65  and  T57.  b.  While  in  monomer  structure  (purple,  PDB                                          ID  4LKP)  Gln65  is  pointing  out  of  toward  the  solvent  and  T64  makes  a  hydrogen  bond                                            to  Thr57  via  water  network.  Res.  57,  64  and  65  are  represented  in  sticks.  Water                                                molecules  are  in  red.................................................................................................108     Figure  III-­‐10:  a.  Size  exclusion  chromatogram  of  Gln65  mutants  of  hFABP5  during  purification  by                                                  SEC.  b.  Size  exclusion  chromatogram  of  Thr64  mutants  of  hFABP5  (Thr63E  pink,                                                  T63L  green)  during  purification  by  SEC..................................................................109     xix   Figure  III-­‐11:  Phase  relationship  in  iLBP  family...........................................................................111     Figure  IV-­‐1:  a.  Polar  residues  around  the  ligand  mutated  in  order  to  remove  polarity  from  the                                              binding  pocket.  Coordinates  obtained  from  PDB  4EXZ  (hCRBPII-­‐Q108K:K40L/retinal).                                                b.  Both  retinal  and  ThioFluor  bind  and  lay  in  the  same  region  of  the  binding                                              pocket.  (the  Thioflour  coordinates  are  from    Q108K:K40L:  T51V:T53S:  R58Y                                              structure,  will  be  discussed  later  in  the  IV-­‐3  section.  c.  The  chemical  Structure  of  the                                              ThioFluor  ,  which    was  synthesized  by  Dr.  Santos....................................................129     Figure  IV-­‐2:    a.  Water  network  hydrogen  bonding  between  T51  and  T53.  b.  Water  mediated                                                                                                                  hydrogen  bonding  between  Q38  and  Q128.  (PDB  4EXZ,  hCRBPII-­‐                                                Q108K:K40L/retinal,  Retinal  is  represented  in  green  sticks)...................................130     Figure  IV-­‐3:  a.  Crystal  structure  of  Q108K:K40L:T51V:T53S:R58W/ThioFluor  and  space-­‐filling                                                            representation  obtained  by  Dr.  Nosrati.  R58W  Is  packed  with  the  N,N-­‐dimethyl                                              amino  moiety  of  the  ThioFlour  tail.  b.  The  water  mediated  hydrogen  bonding                                                between  T53S  and  R58W  is  highlighted..................................................................133     Figure  IV-­‐4:    a.  Overlay  of  Crystal  structures  of  Q108K:K40L:T51V:T53S:R58W/ThioFluor  (cyan)                                                  with  Q108K:K40L:T51V:T53S:R58Y/ThioFluor  (green).  b.  It  shows  Water  mediated                                                hydrogen  bonding  between  R58Y  and    T29............................................................134     Figure  IV-­‐5:    Overlay  of  Crystal  structures  of  Q108K:K40L:T51V:T53S:R58W/ThioFluor  (cyan)                                                    with  Q108K:K40L:T51V:T53S:R58Y/ThioFluor  (green)  and                                                    Q108K:K40L:T53A:R58F/ThioFluor    (salmon)........................................................135     Figure  IV-­‐6:  The  water  mediated  hydrogen  bonding  between  Tyr19  to  the  thiophene  sulfur  of                                                ThioFluor  highlighted  in  the  Crystal  structure  of                                                Q108K:K40L:T51V:T53S:R58W/ThioFluor  obtained  by  Dr.  Nosrati.  ......................  136     Figure  IV-­‐7:  a.  Crystal  structure  of  Q108K:K40L:T51V:T53S:R58W:Y19W/ThioFluor.  b.  Water                                              network  is  abolished  in  this  structure.    ThioFluor  and  residues  around  it  are                                              represented  sticks....................................................................................................137     Figure  IV-­‐8:  Crystal  structure  of  Q108K:K40L:T51V:T53S:R58W:Y19W:A33W:L117E/ThioFluor                                              obtained  by  Alireza.    ThioFluor  in  cyan  and  residues  around  it  in  green  are    shown  in                                              sticks........................................................................................................................138     Figure  IV-­‐9:  Mutated  residues  which  discussed  in  this  chapter  (show  in  blue)  in  an  attempt  to                                                encapsulate  the  binding  cavity.  Crystal  structure  is  of                                                Q108K:K40L:T51V:T53S:R58W:Y19W/ThioFluor.....................................................139     Figure  IV-­‐10:  Crystal  structure  of  Q108K:K40L:T51V:T53S:R58W:Y19W:A33W:L117E/ThioFluor.                                                    ThioFluor  and  highlighted  residues  are  represented  as  sticks...............................140   xx   Figure  V-­‐1:  Schematic  view  of  the  BE  isozymes  and  SS  isozymes  in  the  amylopectin  cluster....159   Figure  V-­‐2:  Overall  structure  of  rice  Branching  Enzyme  I  in  complex  with  maltododecaose                                                                                  (M12).  The  N-­‐terminal  domain  is  shown  in  green,  CBM48,  pink,  center  catalytic                                                                      domain,  cyan,  and  the  C-­‐terminal  ι-­‐amylase  C  domain  in  blue.  Carbohydrates  are                                            represented  in  sticks  and  colored  by  atom  type,  carbons  in  yellow  and  oxygens  in                                            red.  One  oligosaccharide,  M12,  binds  exclusively  into  the  catalytic  domain  and  hangs                                            over  the  catalytic  groove  without  reaching  inside  (site  II),  while  the  five  glucose  units                                            visible  for  the  second  molecules  (site  I),  between  three  domains:  the  N-­‐terminal,  the                                            carbohydrate  binding  module,  and  the  catalytic  domain.  .......................................161     Figure  V-­‐3:  Surface  depiction  of  RBEI  in  complex  with  M12.  At  the  center  of  this  groove,                                                residues  involved  in  catalysis  are  shown  in  blue.  These  catalytic  residues,  Y235,                                              D270,  H275,  R342,  D344,  E399,  H467  and  D468  according  to  RBEI  sequence                                              numbering,  (2)  were  predicted  based  on  biochemical  and  structural  data  of  ι-­‐                                            amylase.  (3,  4)  ............................................................................................................162     Figure  V-­‐4:  Binding  site  I:  detailed  interactions  between  the  oligosaccharide  and  RBEI.  The                                              protein  atoms  are  colored  by  type:  C  in  blue  marine,  O  in  red  and  N  in  dark  blue.                                            M12  atoms  are  also  colored  by  type  with  C  in  yellow  and  O  in  red.  Glucose  units  are                                            numbered  in  red.  Hydrogen  bonds  are  shown  in  dotted  black  lines.  Water  molecules                                            interacting  in  this  site  are  represented  in  spheres  and  colored  in  cyan...................163     Figure  V-­‐5:    Binding  site  II,  detailed  interactions  between  the  oligosaccharide  and  RBEI.                                                Numbers  for  glucose  units,  and  the  protein  atoms  are  colored.  Hydrogen  bonds  and                                                water  molecules  are  also  represented.  Hydrogen  bonds  are  shown  in  dotted  black                                                  lines.  .......................................................................................................................164     Figure  V-­‐6:  Overall  structure  of  RBEI.  Active  site  residues  are  shown  in  blue  sticks.  Two                                            oligosaccharide  molecules  are  in  yellow  sticks.  Oxygen  atoms  are  in  red.  So  far,  we                                            have  not  identified  any  glucose  units  close  and  around  the  active  site  yet.............166     Figure  V-­‐7:  Absorbance  versus  Time  for  the  D344A  mutation  of  truncated  RBEI.The  graph  is  the     average  for  three  trials.    Slope/Protein  indicates  the  activity  (U/mg)  of  the                         enzyme......................................................................................................................167     Figure  V-­‐8:  Detailed  interactions  between  the  Glu7  in  binding  site  II  and  residue  E534  of  RBEI.                                            All  interactions  between  them  are  in  between  2-­‐3.5  Å.    Numbers  for  glucose  units,                                                  and  the  protein  atoms  are  colored.  Hydrogen  bonds  and  water  molecules  are  also                                              represented.  Hydrogen  bonds  are  shown  in  dotted  black  lines.  .............................168     Figure  V-­‐9:    Absorbance  versus  Time  for  the  E534A  mutation  of  truncated  RBEI:  The  graph  is  the                                                average  for  three  trials.  The  decrease  in  absorbance  is  linear  for  the  first  5  min  then                                                reaches  a  plateau  due  to  the  high  activity  of  the  enzyme......................................169   xxi   KEY  TO  SYMBOLS  AND  ABBREVIATIONS   Å     Angstrom   PDB  ID     Protein  Data  Bank  Identifier   PSB     Protonated  Schiff  Base   SB     Schiff  Base   iLBP       intracellular  Lipid  Binding  Proteins   hCRBPII     human  Cellular  Retinol  Binding  ProteinII   hCRABPII     human  Cellular  Retinoic  Acid  Binding  ProteinII   hFABP4   human  Fatty  Acid  Binding  Protein4   hFABP5   human  Fatty  Acid  Binding  Protein5   σ     Sigma   UV     Ultra  Violet   WT       Wild-­‐Type   IPTG     Isopropyl  &-­‐D-­‐1-­‐thiogalactopyranoside   PCR     Polymerase  Chain  Reaction   PEG     Polyethylene  glycol   Vis     Visible   Kd     Dissociation  constant   SDS-­‐PAGE   Sodium  dodecyl  sulfate  polyacrylamide  gel  electrophoresis   KDa   Kilo  Dalton       xxii   E.  Coli's   Escherichia  coli   min     minute   S     Second   Îľ     Extinction  coefficient   QY     Quantum  Yield   Îťmax     Maximum  wavelength   Îťex     Excitation  wavelength   Îťem     Emission  Wavelength   mM     Milimolar     ÎźM       Micromolar     nM       Nanomolar     mol       mole   mmol       millimole   mg       milligram   ml       milliliter   DNA     deoxyribonucleic  acid   dNTP       deoxynucleotide  triphosphates     rpm       rotation  per  minute   °C       degrees  of  centigrade   K       degrees  of  kelvin   pH       Logarithmic  scale  of  hydrogen  ion  activity     NaOH       Sodium  hydroxide   xxiii   Equiv       equivalent   RMSD       root  mean  square  deviation   Ala,  A       Alanine   Arg,  R       Argannine   Asn,  N     Aspargine   Asp,  D       Aspartate   Cys,  C       Cysteine   Gln,  Q       Glutamine   Glu,  E       Glutamate   His,  H       Histitidine   Ile,  I       Isoleucine   Leu,  L       Leucine   Lys,  K       Lysine   Met,  M     Methionine   Phe,  F       Phenylalanine   Pro,  P       Proline   Ser,  S       Serine   Thr,  T       Threonine   Trp,  W     Tryptophan   Tyr,  Y       Tyrosine   Val,  V     Valine   xxiv   DS                         Domain  Swapping     LB     Luria  Broth       R-­‐factor     reliability  factor     RT     Room  temperature     RMSD       root-­‐mean-­‐square  deviation                                   xxv   Chapter  I:  Structural  Studies  of  Human  Cellular  Retinol  Binding  Protein   II  (hCRBPII)  bound  to  Retinol  and  Retinal.             I-­‐1  Introduction     Vitamin   A   (retinol)   is   an   essential   micronutrient   that   plays   a   key   role   in   vision,   cell   growth,   differentiation   metabolism   etc.   (5)   Because   of   its   low   solubility   in   aqueous   medium,   retinol  is  bound  by  specific  binding  proteins  in  body  fluids  and  within  the  cell  for  its  transport,   bioavailability   and   stability.   Retinol   is   converted   to   retinal,   which   is   the   essential   molecule   of   vision,  and  also  to  retinoic  acid,  which  is  essential  in  cell  differentiation.  (6)   In  different  tissues,  different  binding  proteins  play  important  roles  in  transport,  storage   and   metabolism   of   vitamin   A,   such   as   Retinol   Binding   Proteins   (RBP),   Cellular   Retinoic   Acid   Binding   Proteins   (CRABP),   Cellular   Retinol   Binding   Protein   (CRBP),   Cellular   Retinaldehyde     Binding   Protein   (CRALBP)   and   Interphotoreceptor   Retinol   Binding   Protein   (IRBP). In   all   those   proteins,  the  hydrophobic  ligand  and  protein  have  non-­‐covalent  interactions.(7,  8)   CRBPs   belong   to   the   family   of   intracellular   lipid   binding   proteins   (iLBP’s).     iLBPs   are   a   sub-­‐family  of  the  large  family  of  calycins  that,  in  addition  to  the  iLBP’s,  include  the  avidins  and   lipocallins.    All  the  Calycins  have  a  beta-­‐sandwich  structure  in  common  and  all  are  chaperones   for   their   hydrophobic   ligands,   which   bind   in   their   relatively   large   binding   cavities,   though   sequence  homology  between  the  subfamilies  is  less  than  10%.   (6,  9-­‐13)  ILBPs  are  relatively  small   proteins   (126-­‐140   amino   acid)   found   in   the   cytosols   of   fish,   amphibians,   reptiles,   birds   and   mammals.   They   responsible   for   shuttling   insoluble   hydrophobic   molecules,   including   retinal,   retinoic   acid,   various   long   chain   fatty   acids,   eicosinoids,   cholesterol   and   hemes,   throughout   the   1   cell.(6,   10-­‐13)   Their   structure   consists   of   a   ten-­‐stranded  𝛽 -­‐barrel,   and   two   short  𝛼-­‐helices   that   cover  the  binding  pocket,  with  the  ligand  buried  deeply  within  the  binding  site.  Basically,  two   short  ι  helixes  cover  the  binding  pocket  like  a  lid,  isolating  the  ligand  from  bulk  solution  and  the   hydrophobic  ligand  is  deeply  buried  inside  the  binding  pocket.   (14)(Figure   I-­‐1a)   This  family  also   includes  the  Cellular  Retinoic  Acid  Binding  Proteins  (CRABP),  and  the  Fatty  Acid  Binding  Proteins   (FABP)  with  quite  similar  structures.(8)     b. a.             Figure   I-­‐1:   a.   An   overlay   of   the   structures   of   hCRBPII   (green   and   cyan),   hCRBPIII  (1GGL,  salmon),  hCRBPIV  (1LPJ,  yellow)  and  hCRBPI  (1KGL,  pink)   b.   An   overlay   of   the   structures   of   apo   hCRBPII   (2RCQ,   gray),   wt   hCRBPII   bound   to   all-­‐trans-­‐retinal   (4QYP,   green)   and   to   all-­‐trans-­‐retinol   (4QYN,   cyan).   There  are  four  identified  isoforms  for  CRBP  (CRBPI,  II,  III  and  IV),  among  them  only  CRBPI   and  II  have  been  shown  to  be  involved  in  the  metabolism  and  transport  of  intracellular  retinol   so  far.  HCRBPI  is  found  predominantly  in  the  liver  and  kidney  while  hCRBPII  is  more  prevalent  in   the   small   intestine.   (5,   15,   16)It   was   reported   that   hCRBPII   regulates   retinoid   metabolism   in   intestine  and  facilitates  the  reduction  of  retinal  to  retinol  and  the  subsequent  esterification  of   retinal   to   retinyl   ester.   HCRBPIII   mRNA   is   expressed   in   kidney   and   liver   thus   suggesting   an   impotent   role   as   an   intracellular   mediator   of   retinol   metabolism   in   these   tissues.     HCRBPIV   2   mRNA   is   expressed   primarily   in   heart,   and   transverse   colon.   However,   physiological   roles   of   newly  identified  hCRBPIII  and  IV  are  still  unknown.(8,  17)               Figure   I-­‐2:   The   multiple   sequence   alignment   between   human,   rat   CRBPs   and  zebra  fish  CRBP.     CRBPI  and  II  share  56%  sequence  identity  and  both  proteins  can  bind  all-­‐trans-­‐retinol,  all   trans-­‐retinal  and  13-­‐cis-­‐retinol,  while  neither  bind  9-­‐cis-­‐retinol  though  CRBPII  is  found  to  be  the   more  selective  for  all-­‐trans-­‐retinoids.  (18-­‐20)  (Figure  I-­‐2)  As  the  first  crystal  structures  of  CRBP,  we   can  point  to  the  apo  and  all-­‐trans-­‐retinol  bound  structure  of  rat  CRBPII,  which  was  published  in   1993.(21-­‐23)   Comparison   of   the   two   structures   showed   that   no   significant   differences   were   observed  upon  ligand  binding.  Later  on,  the  crystal  structures  of  human  CRBPII  and  a  zebra  fish   CRBP,   both   bound   with   all-­‐trans-­‐retinol,   were   also   reported.   (21,  22)However,   in   both   cases   the   electron   density   of   the   end   of   the   chromophore   was   not   clear.   The   ligand   was   modeled   as   a   mixture  of  retinol,  and  a  retinol  derivative  whereas  the  C13-­‐C14  bond  is  a  single  bond.  (2RCT,   1KQW)   In   our   lab   and   in   collaboration   with   prof.   Borhan’s   group,   we   were   the   first   group   to   achieve   the   structure   of   an   all-­‐trans-­‐retinal   bound   CRBP   (hCRBPII)   by   Dr.   Nossoni.     We   also   3   were   the   first   to   show   that   a   retinoid   derivative   virtually   identical   to   that   seen   in   previously   published   CRBPII   structures   can   be   generated   by   X-­‐ray   radiation   damage,   and   this   process   is   dependent  on  the  wavelength  of  the  X-­‐ray  radiation.  In  addition,  we  finally  determined  the  first   bonafide  structure  of  retinol-­‐bound  hCRBPII.(6)    (Figure  I-­‐1b)   I-­‐2  The  affinity  of  hCRBPII  for  retinol  and  retinal    The   affinity   of   rat   CRBPII   for   both   retinol   and   retinal   has   been   measured   using   both   a   fluorescence   quenching   assay   and   Fluorine   NMR.   The   most   recent   study   showed   rat   CRBPII   binds   to   retinol   and   retinal   with   similar   affinity   (a   Kd   of   about   10   nM).(20,  24,  25)   As   far   as   we   knew,   no   affinity   measurements   for   hCRBPII   had   been   reported,   and   then,   we   decided   to   measure   its   affinity  to  both  retinal  and  retinol  using  a  Tryptophan  fluorescence-­‐quenching  assay    (Described   in  Experimental  Section).    While  the  binding  of  retinal  was  significantly  weaker  compared  with   that   most   recently   reported   for   the   rat   protein,   130Âą10   nM   VS   10nM   in   rat   CRBPII,   Figure   I-­‐ 3d)(24),  hCRBPII  showed  significantly  higher  affinity  for  retinol,  with  a  lower  limit  of  2.2Âą7.9  nM   (Figure   I-­‐3a).   The   binding   constant   appeared   to   be   too   high   for   accurate   measurement   using   our  fluorescence  assay.    This  seems  to  indicate  a  potential  difference  in  the  interaction  of  rat   versus  human  CRBP  for  these  ligands.  The  ligand  is  stabilized  within  binding  pocket,  allowing  it   to  preserve  its  yellow  colour  at  room  temperature  and  in  light  for  more  than  three  days,  while   in  ethanol  solution  a  quarter  of  the  chromophore  is  degraded  after  4  hours  when  exposed  to   air.   4     a   Binding  A ffinity  of   WT  hCRBPII  to  A ll-­‐ Trans-­‐Retinol   b     Kd=  2.2nMÂą7.9     Binding  A ffinity  of   T51I-­‐hCRBPII  to  All-­‐ Trans-­‐Retinol   Kd=  2.2nMÂą6.7         c     Binding  A ffinity  of   T51I-­‐hCRBPII  to  All-­‐ Trans-­‐Retinal   Kd=46nM Âą  10       d   Binding  A ffinity  of  WT   hCRBPII  to  All-­‐Trans-­‐ Retinal   Kd=  130nM Âą  14         Figure   I-­‐3:   Ligand   binding   measurements   using   the   fluorescence   quenching   of   tryptophan   (excitation   at   280   nm,   emission   at   350   nm).   a.   Titration   of   WT-­‐hCRBPII   with   all-­‐trans-­‐retinol.   b.   Titration   of   hCRBPII   T51I   mutant   with   all-­‐trans-­‐retinol.   c.   Titration   of   hCRBPII   T51I   mutant   with   all-­‐trans-­‐retinal.     d.   Titration   of   WT-­‐hCRBPII   with  all-­‐trans-­‐retinal.         I-­‐3  Structure  of  the  retinal-­‐bound  hCRBPII  complex   Retinal-­‐bound   hCRBPII   crystallizes   in   the   P1   space   group   with   four   molecules   per   asymmetric   unit   (Described   in   experimental   section,   Table   I-­‐1).(6)   Electron   density   corresponding  to  the  entire  ligand  was  very  clear  in  only  one  of  the  four  chains  (chain  B).    In   chain   C,   density   for   retinal   is   well   defined   for   all   atoms   but   C15   and   the   hydroxyl   group.     In   molecule  A,  electron  density  is  relatively  well-­‐defined  for  the  carbonyl  group  and  C11-­‐C15,  but   is   weaker   for   the   rest   of   the   molecule.   Scattered,   uninterruptable   electron   density   is   seen   in   5   molecule   D.   The   overall   structure   of   the   protein   is   little   changed   either   in   the   apo   state,   or   when  bound  to  either  retinal  or  retinol.  (Figure  I-­‐1a,  1b)  The  retinal  is  essentially  identical  in  all   three  molecules  where  the  electron  density  is  clear  (Figure  I-­‐4a).     The  β-­‐ionone  ring  adopts  the  6-­‐s-­‐trans  conformation  in  both  ligands.  Two  hydrophobic   residues,   Phe16   and   Leu77,   surround   the   ionone   ring   of   the   ligand,   locking   the   chromophore   in   position   (Figure  I-­‐4b,  4c).   The   distances   between   the   C5   methyl   group   of   retinal   and   Leu77   and   Phe16  are  about  3.7Å  and  3.8Å  respectively.  (Figure  I-­‐4b)       Also,  the  carbonyl  group  of  retinal  makes  a  water  mediated  interaction  with  Gln108  and   Lys40.  The  side  chain  of  Gln108  is  fixed  in  position  by  a  small  water  network,  including  the  main   chain  carbonyls  of  Thr1  and  Asp91,  and  the  side  chains  of  Gln4,  and  Gln108  (Figure  I-­‐4a).  This   network   is   unique   to   hCRBPII   and   is   not   seen   in   zebra   fish   CRBP,   rat   CRBPII   and   hCRBPI   (PDB   entry  1KGL),  hCRBPIII  (PDB  entry  1GGL)  and  hCRBPIV  (PDB  entry  1LPJ).  (14)                       6         a.   C14         C15 C13 C12 C11 b.               c.               Figure   I-­‐4:  Retinal  binding  in  hCRBPII.  a.  An  overlay  of   retinal  from  mol  A  (cyan),  mol  B   (blue)   and   Mol  C  (pink),  showing  the  similarity  in  binding  in  the  three.     b.  Stick  representation  of   the  interaction  between    Phe16,  Leu77  (both  with  green  carbons,  N  blue  and  O,  red)   and  the   retinal  ionone  ring  (blue  carbons,  Mol  B).  c.  Space  filling  depiction  of  b.     7   In   this   case,   water   must   act   as   hydrogen   bond   donor   to   the   two   main-­‐chain   carbonyl   groups   (Thr1  and  Asp91),  only  the  lone  pairs  on  the  oxygen  of  water  are  left  to  have  hydrogen  bonding   to   the   amide   nitrogen   of   Gln4,   defining   its   orientation.   Therefore   it   is   the   carbonyl   oxygen   of   Gln4   that   makes   a   hydrogen   bond,   necessarily   with   the   amide   of   Gln108,   defining   its   orientation   as   well.   The   orientation   of   Gln108,   with   the   carbonyl   group   pointing   toward   the   carbonyl   of   the   retinal,   abolishing   the   direct   hydrogen   bond   to   the   retinal   carbonyl   oxygen   and   necessitates  the  water-­‐mediated  hydrogen  bond.  The  hydrogen  bond  between  this  water  and   the  retinal  carbonyl  is  2.4Å,  which  is  an  ideal  distance  for  a  strong  low-­‐barrier  hydrogen  bond   (Figure   I-­‐5   and   Figure   I-­‐9c).    This  water  makes  an  additional  hydrogen  bond  with  the  ξ-­‐amino   group  of  Lys40,  and  is  fixed  in  space  by  two  hydrogen  bonds  in  the  binding  pocket.       Retinal Gln4 3.0Å   2.4Å Thr1 2.5Å 2.5Å   2.8Å 2.9Å Gln108 2.7Å Asp91 Lys40     Figure  I-­‐5:  The  water  mediated  interaction  of  the  retinal  hydroxyl   group   in   hCRBPII.     Atoms   colored   by   type,   hydrogen   bonded   distances  are  shown.       I-­‐4  Wavelength-­‐dependent  damage  of  the  retinol  in  hCRBPII   As  previously  mentioned,  the  structures  of  the  CRBPII-­‐retinol  complex  from  rat,  human   and  zebra  fish  have  already  been  determined.  But  the  presence  of  the  retinoid  derivative  that   8   also   occupies   the   binding   site   in   human   and   zebra   fish   structures   made   our   interpretation   complicated.  (21,  22)     a       b     c   d       e   f       Figure  I-­‐6:     a.   The  contoured  2Fo-­‐Fc   electron   density   map   at   1.0   σ  for   rearranged   retinol   from   data   collected   at   11   KeV/the   highest   flux   (4QYN,   blue).   b.   The     chemical   structure   of   a   possible   retinoid   derivative   consistent   with   the   crystallographic  data.  c.  The  overlay  of  bonafide  retinol  structure  (chain  C)  from   data   collected   with   an  attenuated   X-­‐ray  beam   at   an   energy   of   7   KeV   (4QZT,   Pink)     and   the  structure  of   rearraged  retinoid  derivative   (chain  B)  from  data   collected  at   11  KeV  and  the  highest  flux  (4QYN,  blue).  d.  The  structures  of  the  noncanonical     retinoid   (refined   in   two   conformations)   previously   seen   in   the   human   CRBPII   (2RCT,  yellow),  and  our  current  rearranged   structure  (4QYN,  blue)   were  overlaid.     The   structures   of   the   canonical   and   noncanonical   retinoid   previously   seen   in   e.   the  zebrafish  CRBP  (1KQW,  green),  and  our  current  rearranged  structure  (4QYN,   blue)   were   overlaid.  f.  The  overlaid   structures  of  the   canonical  retinol  obtained  in     the  rat  CRBPII  (1OPB,  purple)  and  our  current  bonafide  all-­‐trans-­‐retinol  structure   in  human  CRBPII  (4QZT,  pink).     This   other   retinoid   species,   postulated   to   be   a   degradation   product   from   the   original   retinol,   is   characterized   by   a   torsion   angle   about   the   C13-­‐C14   bond   that   is   around   60°,   inconsistent   with   the   trans   double   bond   between   these   two   carbons   expected   for   retinol.   (Figure   I-­‐6b,   9a   and   9b)   The   hydroxyl   group   of   this   species   makes   hydrogen   bonds   to   both   9   Gln108  and  Lys40,  essentially  occupying  the  same  position  as  the  water  molecule  responsible   for   the   water   mediated   interaction   seen   in   our   hCRBPII-­‐retinal   complex.   In   rat   CRBPII-­‐retinol   complex   shows   only   retinol   in   the   binding   pocket,   with   the   hydroxyl   group   in   a   distinctly   different  position  to  that  seen  in  the  other  structures,  allowing  it  to  make  hydrogen  bonds  to   both  the  amide  oxygen  and  nitrogen  atoms  of  Gln108.  (Figure  I-­‐7c  and  9c)  Since  we  were  not   confident   on   the   previous   hCRBPII-­‐retinol   complexes,   the   structure   of   the   hCRBPII-­‐retinol   complex  was  re-­‐determined  in  a  new  P1  crystal  form  with  2  molecules  per  asymmetric  unit  in   our  group.    (Table  II-­‐1)     b   a           c   (a)   Asn91 (b) Thr1 Gln108 Gln4 W   W retinol Asp91 Val110     Gln4 Gln108 retinol   I-­‐7:     a.   The   rearranged   Figure   retinol   from   data   collected   at   11   (d) KeV   and   the   highest   flux   (c) Phe4 Phe4 Thr51 (4QYN,  blue)  and  the  interaction  of  alcohol  moiety  with  the  neighboring  residues  of  Gln108   Ile51 and    Lys40.   b.   The   structures   of   the   retinal   bound   hCRBPII   (4QYP,   cyan,   chain   B),   and   our   current   rearranged   retinol   structure   (4QYN,   blue)   were   overlaid.   The   ordered   water   Asp91 molecule  (W)  seen   in  all-­‐trans-­‐retinal  crystal  structure,  is  occupied  by  the  hydroxyl  group   of   the  noncanonical  retinoid  (indicated  in  dashed  circle).  Similar  case  in  retinol  structure  from   data   collected   at   7   KeV   and   an   attenuated   beam   Gln108 has   been   seen.   c.   Interactions   around   Gln108 retinol and retinoid retinol Gln108  of  retinol  bound  rat  CRBPII  (1OPB,  purple).   10   Asp91 Interestingly,  The  same  retinoid  derivative  was  seen  in  our  structure.  (Figure   I-­‐6a,   6d   and   6e)   The  hydroxyl  of  the  retinoid  derivative  makes  hydrogen  bonds  to  both  Gln108  and  Lys40,  and   occupies   a   position   that   is   essentially   identical   to   that   of   the   water   in   the   retinal   complex   (Figure  I-­‐7).  No  evidence  for  retinol  is  seen  in  our  structures.  Though  the  exact  identity  of  the   retinoid  is  unclear  (Figure  I-­‐6b).  NMR  spectrum  of  the  retinol  sample  was  taken  by  Dr.  Yapici  to   determine  the  purity  of  the  retinol.    The  NMR  showed  the  retinol  to  be  quite  pure  with   no  significant  extraneous  peaks.    Further,  Dr.  Yapici  performed  a  series  of  analytical  HPLC  of  the   sample,   which   they   also   showed   only   a   single   peak   consistent   with   retinol.   (Figure   I-­‐8).   Since   our  crystals  were  all  grown  at  a  relatively  low  pH  of  4.6,  we  investigated  the  possibility  that  the   low   pH   of   the   crystallization   condition   could   lead   to   the   retinoid   derivative   by   incubating   retinol,   both   alone   and   in   the   presence   of   hCRBPII   overnight.   In   both   cases   a   decomposition   product  was  identified  by  HPLC.    However,  this  product  was  less  polar  as  evident  from  the  HPLC   retention   times,   and   is   inconsistent   with   a   hydroxyl-­‐group   containing   compound.   Then   we   considered  that  the  low  pH  product  was  a  dehydration  product,  similar  to  that  previously  seen,   but   was   not   the   compound   seen   in   the   structure   (Figure   I-­‐a,   6b).(26)   We   then   grew   over   30   crystals  of  the  complex,  and  Dr.  Yapici  dissolved  them  in  phosphate  buffered  saline  extracted   the    retinoids  by  hexane  and  analysed  them  by  HPLC  and  UV/Vis  spectroscopy.      all  of  the  other  data  sets,  the  human,  zebrafish,  and  both  of  our  hCRBPII  complex  structures,   were   collected   at   synchrotron   sources   using   about   1Å   (11   KeV)   wavelength   radiation.   (Figure  I-­‐ 6,  9)(23)     11       a b   .   c d .   e   Figure   I-­‐8:   This   experiment   performed   by   Dr.   Yapici   a.   UV-­‐vis   spectrum   of   all-­‐trans-­‐ retinol.     b.     The   HPLC   trace   of   all-­‐trans-­‐retinol.   c.     UV-­‐vis   spectrum   of   the   retinol   extracted   from   the   crystals   and   purified   by   HPLC.  d.   HPLC   chromatogram   of   the   retinol   extracted   from   crystals.   e.     HPLC   chromatogram   of   bonafide   all-­‐trans-­‐retinol   and   the   retinol  extracted  from  crystals.  Showing  that  they  are  the  same  species.   This  showed  only  a  single  peak  that  co-­‐eluted  precisely  with  retinol,  indicating  that  the   crystals   do   in   fact   contain   only   retinol.   (Figure   I-­‐8)   It   therefore   appears   that   the   retinoid   derivative   seen   in   these   structures   is   due   to   X-­‐ray   induced   damage   during   data   collection.   Consistent  with  this  possibility,  It  was  observed  that  the  yellow  colour  of  the  crystal  changed  to   dark   orange   after   its   exposure   to   the   X-­‐ray   beam   and   this   colour   is   maintained   to   the   end   of   12   data   collection.(27-­‐29)   It   is   interesting   to   note   that   for   the   rat   CRBPII   retinol   complex,   the   only   bonafide  retinol-­‐bound  CRBPII  structure,  data  was  collected  using  a  Cu  KÎą  home  X-­‐ray  source,     In  an  effort  to  finally  produce  a  bonafide  hCRBPII  retinol  complex,  we  collected  a  variety   of   data   sets   at   the   synchrotron.     Since   most   X-­‐ray   damage   of   protein   crystals   are   thought   to   be   dosage   related.   We   first   collected   data   at   11   KeV   but   with   the   beam   attenuated   to   less   than   20%  intensity.  The  electron  density  was  still  consistent  with  only  the  rearranged  retinol  in  the   active   site   (data   not   shown).   Then   we   investigated   the   effect   of   X-­‐ray   energy   by   collecting   data   at  7  KeV,  8  KeV  and  9  KeV,  all  using  a  beam  attenuated  to  20%  of  the  original,  and  an  additional   data   set   at   11   KeV.   Dr   Spencer   Anderson   from   Argonne   National   Laboratory   helped   us   in   collecting   the   attenuated   data   at   multiple   wavelengths.   Surprisingly,   the   retinoid   electron   density   in   the   7   KeV   data   set   clearly   showed   an   un-­‐rearranged   all-­‐trans-­‐retinol   (Figure  I-­‐9c).  All   the   other   data   sets   showed   some   mixture   of   probably   retinol   and   the   rearranged   product.   Together,   these   data   show   that   the   rearranged   product   is   a   result   of   both   dosage   and   wavelength-­‐dependent   X-­‐ray   damage,   and   that   data   collection   at   lower   energy   results   in   the   first  structure  of  the  bonafide  hCRBPII-­‐retinol  complex.  (Figure  I-­‐9d,  10,  1b)(6)   I-­‐5  The  structural  difference  between  retinol-­‐bound  hCRBPII  and  retinal-­‐bound.       The  difference  in  binding  of  retinal  versus  retinol  lies  in  the  orientation  of  the  hydroxyl   group   (Figure   I-­‐10c,   10d).   In   the   hCRBPII   retinol   complex   the   hydroxyl   is   positioned   to   make   hydrogen   bonds   with   Lys40   and   Gln108,   making   hydrogen   bonds   with   both   carbonyl   oxygen   and   nitrogen   atoms   of   Gln108   simultaneously   (an   interaction   that   would   be   impossible   for   retinal,  which  cannot  act  as  hydrogen  bond  donor  to  the  amide  carbonyl  oxygen).  This  results  in   13   a  different  orientation  around  C15,  making  room  for  the  water  molecule  between  Gln108  and   Lys40  and  the  creation  of  the  hydrogen  bond  to  this  water  molecule.   all-trans-retinol retinoid moiety ψ = C15-C14 / C13-C12 ψ = C15-C14 / C13-C12   a  (a) ψ1 = 174.4˚ C13   C13 C12 C14 C12 C14 C13 C15 C15   ψ2 = 127.6˚ retinoid moiety all-trans-retinol ψ = C15-C14 / C13-C12 b  (b) ψ = C15-C14 / C13-C12 ψ3 = 157.0˚   C13 ψ4 = 90.1˚ C12 C12 C14 C13   C14 C13 C15 C15 c   all-trans-retinol   ψ = C15-C14 / C13-C12 (c)     ψ5 = 178.3˚ C13 C12 C15     C14 C13 d   all-trans-retinol ψ = C15-C14 / C13-C12 (d) ψ6 = 171.6˚   C13   C15 C15 C12 C13 C14 C15 Figure   I-­‐9:   A   comparison   of   the   retinol   and   retinoid   moieties   found   in   the   crystal   structures   of   human,   zebra   fish   and   rat   CRBPII.   a.   The   retinol   (left)   and   retinoid   (right)   moieties   found   in   the   previously   published   hCRBPII   (2RCT,   yellow).   The   value   of   128°   torsion  angle  (ψ)  about  the  C13-­‐C14  bond  is  inconsistent  with  a  double  bond  (180°).  b.     The   retinol  and   retinoid  moieties  seen  in   the  crystal  structure  of  zebrafish  CRBP  (1KQW,   green).     The   torsion   angle   about   the   C13-­‐C14   bond   is   also   inconsistent   with   a   double   bond.  c.  The  retinol  structure  in  the  rat  CRBPII  (1OPB,  purple).  d.    The  retinol  found  in  the   X-­‐ray  structure  of  hCRBPII  using  data  collected  at  an  energy  of  7  KeV  and  an  attenuated   beam  (4QZT,  pink)  with  an  approximate  180°  torsion  angle  around  C13-­‐C14  bond.   14         By   overlaying   these   two   structures,   it   is   clear   that   the   position   of   the   retinol   hydroxyl   group   in   hCRBPII   would   in   fact   clash   with   this   water   molecule.   The   difference   in   affinity   for   retinol  versus  retinal  may  be  explained  by  their  different  interactions  with  hCRBPII.    However,   the  physiological  relevance  of  this  difference  in  binding  constant  is  unclear.         (a) a   Gln108   3.3 Å b(b) .   3.6 Å   2.3 Å Lys40   (c) c   d .   (d) 2.4 Å 3.6 Å   Gln108 Gln108 W W 2.5 Å   2.5 Å Lys40 Lys40 Figure   I-­‐10:  Retinal  versus  retinol  binding  in  hCRBPII.  a.  The  structure  of  the  retinol-­‐ bound  WT  hCRBPII.   b.  The  2Fo-­‐Fc  electron  density  map  contoured  at  1.0σ  around   retinol  of   7Kev   data.  c.  The  structure  of   all-­‐trans-­‐retinal   bound   to   WT  hCRBPII.   d.   The   structures  of  all-­‐trans-­‐retinal  (green)  and  retinol  (magenta).  The   position  of  an   ordered   water  molecule   (W)  occupied  by  the   hydroxyl  group  of  retinol  is  indicated   in  dashed  circle.   I-­‐6  Ligand  Binding  in  CRBPI  versus  CRBPII   In  rat,  CRBPI  and  II  have  similar  binding  affinities  for  all-­‐trans-­‐retinol.   (20)  An  overlay  of   the  crystal  structures  of  retinal-­‐bound  and  retinol-­‐bound  hCRBPII  and  the  structure  of  hCRBPI   (PDB  entry  1KGL)  and  rat  CRBPI  (PDB  entry  1CRB)  illustrate  that  in  hCRBPII  the  ligand  is  more   than  1Å  deeper  in  the  binding  pocket.(14)  The  identity  of  the  amino  acid  at  position  51  appears   15   to   be   the   deciding   factor   for   ligand   position   (Figure   I-­‐11).   Position   51   is   Isoleucine   in   CRBPI   (both  rat  and  human)  while  it  is  threonine  in  hCRBPII.  By  comparing  the  structures  of  human   and  rat  CRBPI  (PDB  entry  1CRB)(30),  It  can  be  seen  that  retinal  cannot  bind  hCRBPI  in  the  same   way  that  it  does  in  hCRBPII.  This  is  due  to  a  steric  clash  between  the  side  chain  of  isoleucine  and   the  retinal  carbonyl  group.         Gln108 Asp91       Thr1 Retinal Ile51 Lys40 Gln4 Phe4 Thr51   Figure   I-­‐11:   The  structures   of  retinol-­‐bound   rat   CRBPI   (magenta  carbons   and  hydrogen   bonds)   and   retinal-­‐bound   hCRBPII   (green   carbons   and   hydrogen-­‐bonds)   are   overlayed.   Amino  acid   labels   are  colored  similarly  when   different   in   the  two  structures.     Note  the   change  in  the  water  network  introduced  by  Phe3.           The  combination  of  the  addition  of  the  larger  isoleucine  at  position  51,  and  the  altered   position  of  the  ligand  leave  no  room  for  a  water  molecule  between  Gln108  and  Lys40.  In  both   CRBPI   structures   a   hydrogen   bond   is   formed   between   Gln108   and   the   retinol   hydroxyl   group   (Figure   I-­‐11).   Though   there   is   no   structure   of   a   CRBPI   bound   to   retinal,   it   is   clear   that   a   hydrogen  bond  to  Gln108  can  only  be  made  if  the  amide  nitrogen  is  pointing  toward  the  retinal   carbonyl,  because  it  must  be  the  hydrogen  donor.     This  is  opposite  to  the  orientation  of  Gln108  in  hCRBPII,  where  the  orientation  of  Gln108   is  determined  by  Gln4  and  a  water  network  as  described  previously.  However,  in  CRBPI,  there  is   16   a  phenylalanine  at  position  4  (Phe3  in  hCRBPI)  instead  of  glutamine,  which  may  allow  Gln108  to   rotate  its  sidechain,  such  that  the  amide  nitrogen  is  positioned  to  make  a  hydrogen  bond  with   the  aldehyde  carbonyl,  allowing  retinal  to  bind  CRBPI  very  similarly  to  that  of  retinol,  resulting   in  the  similar  affinity  seen  for  the  two  ligands  in  CRBPII.     In   an   attempt   to   understand   the   role   of   the   residue   at   position   51,   the   hCRBPII-­‐T51I   mutant   was   produced   and   its   binding   affinity   for   retinol   and   retinal   measured.   The   dissociation   constants   of   this   mutant   for   both   all-­‐trans-­‐retinol   and   all-­‐trans-­‐retinal   were   similar   to   that   of   the   wild   type   protein   (Figure   I-­‐3b,   3d)   indicating   that   position   51   is   not   responsible   for   the   differences  in  affinity  in  the  two  proteins.  Unfortunately  attempts  at  crystallization  of  hCRBPII-­‐ T51I   bound   to   all-­‐trans-­‐retinal   were   not   successful.   However,   crystals   of   the   retinol-­‐bound   CRBPII-­‐T51V   complex   were   obtained   and   its   structure   determined.     However,   this   structure,   collected   with   an   intense   X-­‐ray   beam   with   an   energy   of   11   KeV,   also   showed   the   rearranged   retinol   product   in   the   active   site,   making   it   difficult   to   determine   the   effect   of   a   hydrophobic   residue  at  position  51  on  bonafide  retinol  binding.   I-­‐7  Conclusion   Based  on  the  crystal  structure  of  holo  wt  hCRBPII  with  all-­‐trans-­‐retinal,  there  are  at  least   three  distinct  binding  modes  for  retinoids  in  CRBPI  and  II.  In  the  human  and  rat  CRBPII  retinol-­‐ bound   structures,   retinol   is   able   to   make   hydrogen   bonds   with   both   the   carbonyl   and   amide   nitrogen  of  the  side  chain  of  Gln108.  This  mode  of  interaction  is  not  feasible  in  retinal,  due  to   the  lack  of  a  hydrogen  bond  donor,  resulting  in  a  conformational  change  of  the  carbonyl  leading   to   a   water-­‐mediated   interaction   between   the   retinal   carbonyl   and   Lys40   and   Gln108.   This   is   consistent   with   the   higher   binding   affinity   for   retinol   relative   to   retinal   in   hCRBPII.   Since   retinol   17   is  able  to  make  three  hydrogen  bonds  to  Gln108  and  Lys40.  Also,  the  presence  of  Ile51  causes  a   translation  of  the  ligand  in  the  CRBPI  binding  pocket,  resulting  in  only  a  single  hydrogen  bond   being  made  between  Gln108  and  the  retinol  hydroxyl  group.   I-­‐8  Experimental   I-­‐8-­‐1  Material  and  Method:  Protein  Expression  and  Purification   The  hCRBPII  gene  was  purchased  from  ATCC  and  cloned  in  pET17b  vector,  with  using  the   NdeI   and   XhoI   restriction   enzyme   in   the   N-­‐   and   C   terminus   respectively.   1   ÎźL   of   the   resulting   plasmid  were  transformed  in  50  ΟL  of  E.coli  DH5Îą  (NovagenÂŽ)  competent  cells.  The  cells  were   incubated   for   30   min   in   ice   and   heat   shocked   at   42   °C   for   28   seconds,   then   450   ÎźL   of   Luria-­‐ Bertani  broth  (LB)  was  added  and  the  cells  were  incubated  at  37°C  for  an  hour.  The  resulting   mixture   were   spread   on   an   LB   agar   plate   with   (ampicillin   100   Îźg/mL)   and   incubated   at   37°C   for   16  hours.  A  single  colony  was  picked  from  the  plate  and  inoculated  in  5  mL  of  LB  media  contains   100   Îźg/mL   ampicillin.   The   cell   culture   was   grown   over   night   (12-­‐16   hours)   at   37°C,   then   the   media   were   centrifuged   at   15000   rpm   for   1   min.   DNA   extraction   and   isolation   from   the   cell   pellet  was  done  according  to  the  manufacturers  instructions  (Promega  Wizard  and  SV  Miniprep   (A1330)  DNA  purification  kit).  The  target  gene  was  transformed  into  BL21(DE3)  (InvitrogenTM)   E.coli  competent  cells  for  protein  expression.  A  single  colony  was  picked  and  inoculated  in  25   mL   of   LB,   containing   100   Îźg/mL   ampicillin   at   37°C   for   12-­‐16   hours.   This   media   was   then   transferred   into   1L   of   LB   media   with   ampicillin   (100   mg/L)   to   grow   large   scale   and   incubated   at   37°C  until  OD600  reached  0.5-­‐1.0.  The  protein  expression  was  induced  with  1  mM  isopropyl  β-­‐D-­‐ 1   thiogalactopyranoside   (IPTG)   (purchased   from   Gold   Biotechnology)   overnight   at   25°C.   The   18   cells   were   harvested   by   centrifugation   at   5000   rpm   for   20   min.   The   harvested   cells   were   resuspended  in  Tris  buffer  (10  mM  Tris,  10  mM  NaCl  pH  8.0,  50mL).  The  suspended  cells  were   lysed  by  sonication  and  the  lysed  cells  were  centrifuged  at  4°C  (14,000  rpm,  20  min).     The   supernatant   was   loaded   on   a   Fast   Q   anion   exchange   resin   (purchased   from   GE   healthcare),  which  was  equilibrated  with  buffer  A  (10  mM  Tris,  10  mM  NaCl  pH  8.0).  Then  after   the  binding,  the  resin  was  washed  three  times  with  50  mL  of  buffer  A.  The  protein  was  eluted   with   40   mL   of   elution   buffer   (10   mM   Tris,   100   mM   NaCl   pH   8.0).   The   eluted   protein   was   desalted   using   a   CentriprepÂŽ   centrifugal   filter   (10kDa   cutoff)   at   2500   rpm.   Protein   was   then   purified  on  a  15Q  anion  exchange  column  using  a  BioLogic  DuoFlow  system  for  the  second  step   of  purification.  The  purity  of  the  protein  was  determined  with  SDS-­‐PAGE.     I-­‐8-­‐2    Kd  Determination  via  Fluorescence  Quenching  Assay   The   dissociation   constant   measurement,   (Kd)   for   both   all-­‐trans-­‐retinol   and   all-­‐trans-­‐ retinal   (purchased   from   Sigma)   with   hCRBPII   was   determined   by   fluorescence   quenching   assay.   All   samples   were   stored   in   glass   containers,   as   many   plastic   containers   will   leach   fluorescent   impurities  into  the  sample.  Salinized  glassware  was  used  to  avoid  loss  of  protein  and  a  change   in  protein  concentration.     The   cuvette   was   allowed   to   sit   with   3   mL   of   a   0.01%   gelatin   containing   PBS   buffer   (4   mM   NaH2PO4,   16   mM   Na2HPO4,   150   mM   NaCl,   pH   7.3)   for   30   minutes.   The   clean   cuvette   was   rinsed  once  with  distilled  water,  and  the  protein  solution  was  added  (3  mL,  20  ΟM).  The  sample   was  excited  at  280  nm  with  an  excitation  slit  width  of  1.5  nm.  The  fluorescence  was  measured   at  the  peak  maximum  (345  nm).  This  was  repeated  three  times  until  a  stable  emission  intensity   at  350  nm  was  reached.  Retinal  was  added  to  the  cuvette  in  varying  amounts  from  a  1.5  mM   19   stock   solution   in   ethanol   maintained   in   the   dark.   Care   was   taken   to   ensure   that   the   EtOH   volume   remained   below   2%.   The   titration   was   complete   when   there   was   no   observable   quenching   of   fluorescence   upon   addition   of   the   chromophore.   This   was   plotted   as   concentration   of   chromophore   versus   relative   fluorescence   intensity.   The   data   were   analyzed   by  nonlinear  square  fit  of  the  equation.  The  Kd  for  each  hCRBPII  was  determined  according  to   the  method  previously  described.(31)   I-­‐8-­‐3  Crystallization  and  Structure  Determination   The   pure   WT   hCRBPII   protein   was   concentrated   to   5-­‐10   mg/mL.   The   complexes   of   protein   with   all-­‐trans-­‐retinol   were   prepared   by   adding   3-­‐4   equivalents   of   retinal   and   retinol   solution  (30  mM  retinol  in  ethanol).  The  final  concentration  of  ethanol  in  protein  solution  was   kept   below   10%   V/V.   The   mixtures   of   protein   and   ligands   were   incubated   at   room   temperature   in  Black  LiteSafeTM  Microcentrifuge  Tubes  for  2  hours.  The  crystal  of  the  complex  of  protein  and   ligands  were  prepared  at  room  temperature  by  hanging  drop  vapor  diffusion  in  Limbro  plates   wrapped  in  aluminum  foil  to  prevent  light-­‐initiated  degradation  of  the  retinal  and  retinol.  The   best  crystals  were  grown  using  Hampton  Research  Screens  (the  best  crystals  growing  in  30-­‐35%   PEG  4000  (Sigma-­‐Aldrich),  0.1  M  sodium  acetate  (Columbus  Chemical  Industry)  pH  4.6-­‐4.8  and   0.1   M   ammonium   acetate   (J.   T.   Baker)   with   1   ÎźL   of   hCRBPII-­‐ligand   complex   and   1   ÎźL   of   mother   liquor).  The  crystals  appeared  after  3  days  and  reached  their  maximum  size  in  one  week.  The   crystals  were  flash  frozen  in  liquid  nitrogen  using  a  cryo-­‐protectant  solution  (30%  PEG  4000,  0.1   M   sodium   acetate   pH   4.6-­‐4.8,   0.1   M   ammonium   acetate,   20%   glycerol)  and   stored   in   a   liquid   nitrogen  dewar.     20   All   of   the   diffraction   data   were   collected   at   beamline   21-­‐ID-­‐D,   LS-­‐CAT   (Argonne   National   Laboratory,   Advanced   Photon   Source,   Chicago,   IL)   using   a   MAR300   detector   and   1.00Å   wavelength   radiation   at   100K.   The   diffraction   data   were   indexed   using   the   HKL2000   software   package.(32)   The   structure   was   solved   by   molecular   replacement   using   the   MOLREP   program   implemented   in   the   ccp4   programing   package   using   human   CRBPII   (Protein   data   bank,   accession  code  2RCQ(22))  as  model.  Initial  electron  density  maps  and  structure  refinement  was   performed  using  REFMAC5  in  the  CCP4  suite.(33,  34)   (35)All  rebuilding  and  placement  of  ordered   water   molecules   was   done   manually   using   COOT.(36)   The   chromophore   was   created   using   j-­‐ ligand  and  manually  fitted  in  the  electron  density  at  the  final  step  of  the  refinement.                               21    Table  I-­‐1:  Data-­‐collection  and  refinement  statistics.           wt  hCRBPII-­‐retinol-­‐ wt  hCRBPII-­‐retinol   7KeV     Space   group   P21   P1   a(Å)     34.65   b(Å)     75.14   c  (Å)   54.65     )   Îą(°   90.00   β(°     )   100.78   Îł(°     )   90.00     54.18     68.44   Total  reflection   306503   Unique  Reflection   87302   Completeness   (%)     99.8(84.0)   Molecules   per     Asymmetric  Unit   2   a   36.48     54.14   68.30   107.72   107.64   97.19   96.94   103.59   103.71   a 50.00-­‐1.19(1.21-­‐1.19)                                      P1   36.41   Resolution  (Å)     wt  hCRBPII-­‐retinol-­‐     11KeV   29.79-­‐1.89(1.94-­‐1.89)   76284   34.661-­‐1.496  (1.51-­‐ a   1.50)   358399   34646   76284   a a a 92.36(86.0)   95.20(88.0)   4   4   a 20.19(2.04)   Average   I/σ         a R   merge(%)   9.1(43.4)               Mosaicity  (°)                                                                        0.486                                                            a   a 12.97(6.23)   a a 26.45(2.41)   a 7.8(32.5)     4.2(59.8)     0.58   0.58   Note:  Values  in  the  parenthesis  refer  to  the  last  resolution  shell.                   22   Table  I-­‐1    (cont’d)   ! Refinement! statistics! Rwork/!Rfree!(%)! !         Bond!Length!(Å)! 0.0297! 0.008! 0.007! Angle!(°)! 2.381! 1.079! 1.066!   B!average!for!main! 9.148! 34.628! chain! B!average!for! 16.410! 45.016! water!and!side!   chain! Number!of!water! 507! 331! molecules! ! ! ! ! ! Ramachandran!plot! 97.10! 96.04! Most!favored!(%)!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!! 2.07! 3.21! Allowed!(%)! 0.83! 2.12! Outliers!(%)! 23.69!     !39.549!   468!   ! !   96.20! 2.28!   2.49!     PDB#! 4QYN! 4QZT!               23   4QZU!                         REFERENCES                         24   REFERENCES         1.     2.     3.     4.     5.     6.     7.     8.     9.     10.     Ong,  D.  E.  (1984)  A  NOVEL  RETINOL-­‐BINDING  PROTEIN  FROM  RAT  -­‐  PURIFICATION  AND   PARTIAL  CHARACTERIZATION,  Journal  of  Biological  Chemistry  259,  1476-­‐1482.   Nossoni,   Z.,   Assar,   Z.,   Yapici,   I.,   Nosrati,   M.,   Wang,   W.,   Berbasova,   T.,   Vasileiou,   C.,   Borhan,   B.,   and   Geiger,   J.   (2014)   Structures   of   holo   wild-­‐type   human   cellular   retinol-­‐ binding  protein  II  (hCRBPII)  bound  to  retinol  and  retinal,  Acta  crystallographica.  Section   D,  Biological  crystallography  70,  3226-­‐3232.   Veerkamp,   J.   H.,   Vankuppevelt,   T.,   Maatman,   R.,   and   Prinsen,   C.   F.   M.   (1993)   STRUCTURAL   AND   FUNCTIONAL-­‐ASPECTS   OF   CYTOSOLIC   FATTY-­‐ACID-­‐BINDING   PROTEINS,  Prostaglandins  Leukotrienes  and  Essential  Fatty  Acids  49,  887-­‐906.   Banaszak,   L.,   Winter,   N.,   Xu,   Z.   H.,   Bernlohr,   D.   A.,   Cowan,   S.,   and   Jones,   T.   A.   (1994)   LIPID-­‐BINDING   PROTEINS   -­‐   A   FAMILY   OF   FATTY-­‐ACID   AND   RETINOID   TRANSPORT   PROTEINS,  Advances  in  Protein  Chemistry,  Vol  45  45,  89-­‐151.   Vaezeslami,   S.,   Mathes,   E.,   Vasilelou,   C.,   Borhan,   B.,   and   Geiger,   J.   H.   (2006)   The   structure   of   apo-­‐wild-­‐type   cellular   retinoic   acid   binding   protein   II   at   1.4   A   and   its   relationship   to   ligand   binding   and   nuclear   translocation,   Journal   of   Molecular   Biology   363,  687-­‐701.   Haunerland,   N.   H.,   and   Spener,   F.   (2004)   Fatty   acid-­‐binding   proteins-­‐-­‐insights   from   genetic  manipulations,  Progress  in  lipid  research  43,  328-­‐349.   Lucke,   C.,   Huang,   S.,   Rademacher,   M.,   and   Ruterjans,   H.   (2002)   New   insights   into   intracellular   lipid   binding   proteins:   The   role   of   buried   water,   Protein   science   :   a   publication  of  the  Protein  Society  11,  2382-­‐2392.   Schaap,   F.   G.,   van   der   Vusse,   G.   J.,   and   Glatz,   J.   F.   (2002)   Evolution   of   the   family   of   intracellular   lipid   binding   proteins   in   vertebrates,   Molecular   and   cellular   biochemistry   239,  69-­‐77.   Smathers,   R.   L.,   and   Petersen,   D.   R.   (2011)   The   human   fatty   acid-­‐binding   protein   family:   evolutionary  divergences  and  functions,  Human  genomics  5,  170-­‐191.   Franzoni,   L.,   Lucke,   C.,   Perez,   C.,   Cavazzini,   D.,   Rademacher,   M.,   Ludwig,   C.,   Spisni,   A.,   Rossi,  G.  L.,  and  Ruterjans,  H.  (2002)  Structure  and  backbone  dynamics  of  apo-­‐  and  holo-­‐ cellular   retinol-­‐binding   protein   in   solution,   Journal   of   Biological   Chemistry   277,   21983-­‐ 21997.   25   11.     12.     13.     14.     15.     16.     17.     18.     19.     20.     Crow,   J.   A.,   and   Ong,   D.   E.   (1985)   CELL-­‐SPECIFIC   IMMUNOHISTOCHEMICAL   LOCALIZATION   OF   A   CELLULAR   RETINOL-­‐BINDING   PROTEIN   (TYPE-­‐2)   IN   THE   SMALL-­‐ INTESTINE  OF  RAT,  Proceedings  of  the  National  Academy  of  Sciences  of  the  United  States   of  America  82,  4707-­‐4711.   Vogel,  S.,  Mendelsohn,  C.  L.,  Mertz,  J.  R.,  Piantedosi,  R.,  Waldburger,  C.,  Gottesman,  M.   E.,  and  Blaner,  W.  S.  (2001)  Characterization  of  a  new  member  of  the  fatty  acid-­‐binding   protein   family   that   binds   all-­‐trans-­‐retinol,   Journal   of   Biological   Chemistry   276,   1353-­‐ 1360.   Folli,   C.,   Calderone,   V.,   Ottonello,   S.,   Bolchi,   A.,   Zanotti,   G.,   Stoppini,   M.,   and   Berni,   R.   (2001)   Identification,   retinoid   binding,   and   x-­‐ray   analysis   of   a   human   retinol-­‐binding   protein,   Proceedings   of   the   National   Academy   of   Sciences   of   the   United   States   of   America  98,  3710-­‐3715.   Li,  E.,  Demmer,  L.  A.,  Sweetser,  D.  A.,  Ong,  D.  E.,  and  Gordon,  J.  I.  (1986)  ISOLATION  OF  A   CLONED   CDNA-­‐ENCODING   AN   ABUNDANT   INTESTINAL   RETINOL   BINDING-­‐PROTEIN,   Clinical  Research  34,  A685-­‐A685.   Li,   E.,   and   Norris,   A.   W.   (1996)   Structure/function   of   cytoplasmic   vitamin   A-­‐binding   proteins,  Annual  Review  of  Nutrition  16,  205-­‐234.   Li,   E.,   Qian,   S.   J.,   Winter,   N.   S.,   Davignon,   A.,   Levin,   M.   S.,   and   Gordon,   J.   I.   (1991)   FLUORINE   NUCLEAR-­‐MAGNETIC-­‐RESONANCE   ANALYSIS   OF   THE   LIGAND-­‐BINDING   PROPERTIES   OF   2   HOMOLOGOUS   RAT   CELLULAR   RETINOL-­‐BINDING   PROTEINS   EXPRESSED  IN  ESCHERICHIA-­‐COLI,  Journal  of  Biological  Chemistry  266,  3622-­‐3629.   Calderone,   V.,   Folli,   C.,   Marchesani,   A.,   Berni,   R.,   and   Zanotti,   G.   (2002)   Identification   and   structural   analysis   of   a   zebrafish   apo   and   holo   cellular   retinol-­‐binding   protein,   Journal  of  Molecular  Biology  321,  527-­‐535.   Tarter,  M.,  Capaldi,  S.,  Carrizo,  M.  E.,  Ambrosi,  E.,  Perduca,  M.,  and  Monaco,  H.  L.  (2008)   Crystal   structure   of   human   cellular   retinol-­‐binding   protein   11   to   1.2   angstrom   resolution,  Proteins-­‐Structure  Function  and  Bioinformatics  70,  1626-­‐1630.   Winter,   N.   S.,   Bratt,   J.   M.,   and   Banaszak,   L.   J.   (1993)   CRYSTAL-­‐STRUCTURES   OF   HOLO   AND   APO-­‐CELLULAR   RETINOL-­‐BINDING   PROTEIN-­‐II,   Journal   of   Molecular   Biology   230,   1247-­‐1259.   Kane,  M.  A.,  Bright,  F.  V.,  and  Napoli,  J.  L.  (2011)  Binding  affinities  of  CRBPI  and  CRBPII   for  9-­‐cis-­‐retinoids,  Biochimica  Et  Biophysica  Acta-­‐General  Subjects  1810,  514-­‐518.   26   21.     22.     23.     24.     25.     26.     27.     28.     29.     30.     31.   Malpeli,   G.,   Stoppini,   M.,   Zapponi,   M.   C.,   Folli,   C.,   and   Berni,   R.   (1995)   INTERACTIONS   WITH   RETINOL   AND   RETINOIDS   OF   BOVINE   CELLULAR   RETINOL-­‐BINDING   PROTEIN,   European  Journal  of  Biochemistry  229,  486-­‐493.   McBee,   J.   K.,   Kuksa,   V.,   Alvarez,   R.,   de   Lera,   A.   R.,   Prezhdo,   O.,   Haeseleer,   F.,   Sokal,   I.,   and   Palczewski,   K.   (2000)   Isomerization   of   all-­‐trans-­‐retinol   to   cis-­‐retinols   in   bovine   retinal   pigment   epithelial   cells:   Dependence   on   the   specificity   of   retinoid-­‐binding   proteins,  Biochemistry  39,  11370-­‐11380.   Borshchevskiy,  V.  I.,  Round,  E.  S.,  Popov,  A.  N.,  Bueldt,  G.,  and  Gordeliy,  V.  I.  (2011)  X-­‐ ray-­‐Radiation-­‐Induced   Changes   in   Bacteriorhodopsin   Structure,   Journal   of   Molecular   Biology  409,  813-­‐825.   Matsui,  Y.,  Sakai,  K.,  Murakami,  M.,  Shiro,  Y.,  Adachi,  S.,  Okumura,  H.,  and  Kouyama,  T.   (2002)  Specific  damage  induced  by  X-­‐ray  radiation  and  structural  changes  in  the  primary   photoreaction  of  bacteriorhodopsin,  Journal  of  Molecular  Biology  324,  469-­‐481.   Schreckenbach,   T.,   Walckhoff,   B.,   and   Oesterhelt,   D.   (1977)   STUDIES   ON   RETINAL-­‐ PROTEIN  INTERACTION  IN  BACTERIORHODOPSIN,  European  Journal  of  Biochemistry  76,   499-­‐511.   Cowan,   S.   W.,   Newcomer,   M.   E.,   and   Jones,   T.   A.   (1993)   CRYSTALLOGRAPHIC   STUDIES   ON   A   FAMILY   OF   CELLULAR   LIPOPHILIC   TRANSPORT   PROTEINS   -­‐   REFINEMENT   OF   P2   MYELIN   PROTEIN   AND   THE   STRUCTURE   DETERMINATION   AND   REFINEMENT   OF   CELLULAR   RETINOL-­‐BINDING   PROTEIN   IN   COMPLEX   WITH   ALL-­‐TRANS-­‐RETINOL,   Journal   of  Molecular  Biology  230,  1225-­‐1246.   Vasileiou,   C.,   Vaezeslami,   S.,   Crist,   R.   M.,   Rabago-­‐Smith,   M.,   Geiger,   J.   H.,   and   Borhan,   B.   (2007)   Protein   design:   reengineering   cellular   retinoic   acid   binding   protein   II   into   a   rhodopsin  protein  mimic,  J  Am  Chem  Soc  129,  6140-­‐6148.   Otwinowski,   Z.,   and   Minor,   W.   (1997)   Processing   of   X-­‐ray   diffraction   data   collected   in   oscillation  mode,  Macromolecular  Crystallography,  Pt  A  276,  307-­‐326.   Vagin,   A.,   and   Teplyakov,   A.   (1997)   MOLREP:   an   automated   program   for   molecular   replacement,  Journal  of  Applied  Crystallography  30,  1022-­‐1025.   Vagin,   A.,   and   Teplyakov,   A.   (2010)   Molecular   replacement   with   MOLREP,   Acta   Crystallographica  Section  D-­‐Biological  Crystallography  66,  22-­‐25.   Murshudov,   G.   N.,   Skubak,   P.,   Lebedev,   A.   A.,   Pannu,   N.   S.,   Steiner,   R.   A.,   Nicholls,   R.   A.,   Winn,   M.   D.,   Long,   F.,   and   Vagin,   A.   A.   (2011)   REFMAC5   for   the   refinement   of   macromolecular   crystal   structures,   Acta   Crystallographica   Section   D-­‐Biological   Crystallography  67,  355-­‐367.     27   Chapter  II:  Domain  swapping  in  hCRBPII     II-­‐1  Introduction   Protein   oligomers   have   evolved   because   of   their   advantages   over   their   monomers.   These   advantages   include   the   possibility   of   allosteric   control,   larger   binding   surfaces,   higher   concentration   of   active   sites,   new   active   sites   at   subunit   interfaces   and   higher   stability.   However,  the  mechanisms  for  the  evolution  and  assembly  of  oligomeric  species  during  protein   synthesis   remain   unclear.   Different   mechanisms   have   been   proposed   for   the   evolution   of   protein   oligomers,   three-­‐dimensional   (3D)   domain   swapping   is   among   them.   (37-­‐42)   Three-­‐ dimensional   (3D)   domain   swapping   (DS)   is   a   process   in   which   two   or   more   monomer   protein   molecules   exchange   their   identical   structural   elements   to   form   dimers   or   higher-­‐order   oligomers  (Figure  II-­‐1).  (37-­‐42)         HOOC Closed Interface Hinge Loop Open Interface Domain Swapped   Dimerization H2N   NH2 COOH Monomer H2N   COOH Domain Swapped Dimer Figure   II-­‐1:   Cartoon   illustration   of   3D   domain   swapping.   Identical     structural  units  swap  between  2  monomers,  followed  by  creating  the  open   interface  between  them.   28   The   region   that   differs   markedly   between   monomer   and   oligomer,   called   the   “switch   region”  or  hinge  region”  is  relatively  small.  In  most  cases  it  consists  of  a  loop  in  the  monomer   that  becomes  straightened  in  the  DS  oligomer.  In  many  cases  it  is  this  region  that  seems  to  be   the  most  important  determinant  for  the  occurrence  of  domain  swapping.  It  was  first  discovered   by  Eisenberg  and  coworkers  in  1994  to  account  for  the  dimerization  mechanism  of  diphtheria   toxin  (DT)  (Figure   II-­‐2)  and  was  later  shown  to  be  responsible  for  the  dimerization  of  lyophilized   Bovine   pancreatic   ribonuclease   (Rnase   A,   which   can   have   either   N-­‐terminal   or   C-­‐terminal   swapped  dimers).  So  far  only  about  40  distinctly  different  proteins  have  been  observed  to  form   domain  swapped  dimers.  (43-­‐45)   Diphtheria   toxin   (DT)   is   a   533-­‐residues   protein   toxin   secreted   from   a   bacterium   that   causes  diphtheria.  DT  has  three  domains:  a  catalytic  domain  (C),  a  transmembrane  domain  (T)   and  a  receptor-­‐binding  domain  (R)   (37-­‐42).  The  dimeric  form  of  DT  does  not  form  spontaneously   even  at  higher  concentrations.  However,  DT  can  be  induced  to  form  dimers  by  decreasing  pH,   which   converts   monomeric   DT   into   open   monomers,   which   form   a   DT   dimer   at   high   concentration  by  swapping  the  globular  domain  (R  domain)  (37-­‐42)  The  structure  of  monomer  DT   revealed   the   mechanism   by   which   low   pH   could   trigger   changes   in   monomer   DT   and   thereby   form   an   open   monomer.   In   the   monomer,   the   interdomain   interface   between   the   R   domain   and  C  domain  is  charged,  with  nine  basic  and  three  acidic  residues  on  the  R  domain  interface   and  seven  acidic  residues  on  the  C  domain  interface.  There  are  three  salt  bridges  stabilizing  the   interface   at   neutral   pH.   The   decrease   in   pH   causes   protonation   of   those   acidic   residues   and   disruption   of   the   salt   bridges.   Furthermore,   buried   positive   charges   in   the   interface   favor   the   open   monomer   formation.   During   the   dimerization   of   DT,   higher   order   oligomers   were   29   observed  by  size  exclusion  HPLC.  These  oligomers  include  trimers,  tetramers  and  pentamers  (37-­‐ 42) .  In  most  cases  the  exchanged  region  is  located  at  the  N-­‐  or  C-­‐terminus  of  the  protein,  though   the  exchanged  part  is  found  in  the  middle  of  the  sequence  in  a  few  instances.  (46-­‐49)     a       In  acidic   pH       Monomer    DT   Dimer    DT       b         Figure   II-­‐2:   a.   Left:   DT   monomer   (PDB   ID   4OW6).   Right:   DT   dimer   (4AE1).   The   two     subunits   are  blue   and  pink.  In  acidic  pH  equilibrium   goes  toward  DS  dimer   formation.   b.  Left  structure  of  DS  dimer  of  RnaseA  (PDB  1A2W)  and  cyclic  DS  trimes  (PDB  5SRA).   Domain   swapping,   combined   with   gene   duplication,   has   been   postulated   to   be   responsible  for  the  evolution  of  a  number  of  larger  protein  domains  from  smaller  fragments(50).   Domain   swapping   can   also   lead   to   aggregation,   amyloid   and   fibril   formation,   which   lead   to   30   protein   malfunction(51).   Protein   misfolding   and   aggregation   is   the   cause   of   a   number   of   maladies  including  Parkinson’s  disease,  Alzheimer  disease,  diabetes,  Huntington’s  disease  and   many   others(52-­‐58).     Triggers   of   domain-­‐swapping   include   lyophilization,   non-­‐physiological   pH,   unusually   high   protein   concentration   (59,   60)   (43,   61)   temperature,   mutation   and   even   ligand   binding(62,  63).     The  classic  example  of  domain  swapping  is  bovine  pancreatic  ribonuclease  A  (RNase  A),  which,   the   DS   of   the   N-­‐terminal   exchange,   was   proposed   in   1962   by   Crestfield,   Stein,   and   Moore   to   describe  its  behavior  under  acidic  conditions.  Followed  by  the  first  X-­‐ray  structures  for  a  domain   swapped  dimer  in  the  late  nineties.  In  RNase  case,  either  the  N-­‐terminal  helix  or  the  C-­‐terminal   strand  can  domain  swap.  The  swapped  unit  and  its  oligomeric  state  can  vary  significantly  and   RNase   A   illustrates   the   possibilities   of   domain   swapping   in   the   trimer   both   N-­‐   and   C-­‐terminal   units   are   swapped   and   a   circular   arrangement   succeeds   (Figure   II-­‐2b).   Another   example   is   Cyanovirin-­‐N  (CV-­‐N),  which  is  a  101  amino  acid  cyanobacterial  lectin.  CV-­‐N  inactivates  HIV  and   is   a   general   virucidal   agent   against   other   enveloped   viruses.   The   original   solution   structure   was   a   monomer,   whereas   the   subsequently   determined   X-­‐ray   structures   were   domain   swapped   dimers,  which  each  domain  comprises  a  triple-­‐stranded  β-­‐sheet  with  a  β-­‐hairpin  packed  on  top.   A  helical  linker  is  located  in  the  middle  of  the  sequence.  It  was  shown  that  the  DS  dimer  is  a   kinetically   trapped   folding   intermediate   at   high   protein   concentrations   that   convert   into   the   slightly   more   stable   monomer   form   at   physiological   (>30oC)   temperature.1d,   10   At   room   temperature   or   below,   however,   the   dimer   lifetime   is   sufficiently   long   for   structural   characterization  in  solution.  The  fact  that  both  monomeric  and  domain-­‐swapped  dimeric  CV-­‐N   coexist  in  solution  under  identical  conditions  indicates  that  the  free  energies  of  folding  for  both   31   quaternary   states   must   be   comparable   and   the   kinetic   barrier   between   the   monomer   and   dimer   has   to   be   significant.   This   can   be   altered   by   mutation   of   residues   in   the   hinge   region.   For   instance,   changing   the   proline   in   the   hinge   region   to   glycine   resulted   in   a   substantial   stabilization   of   this   monomeric   P51G   mutant   by   >5   kcal/mol   compared   to   wild-­‐type.   A   S52P   mutant   yielded   predominantly   dimeric   protein   due   to   destabilization   of   the   monomer.   More   and  more  domain  swapped  protein  structures  are  becoming  elucidated,  and,  for  several  cases,   growing   evidence   supports   that   the   dimer   or   multimer   is   an   active,   biological   important   structure.   Regardless   of   whether   domain   swapping   is   a   specific   mechanism   for   regulation   in   vivo,   it   is   becoming   clear   that   domain   swapping   is   a   means   by   which   stable   oligomers   can   be   generated  under  evolutionary  force.  By  protein  expression  or  refolding  at  high  concentrations,   suggesting   that   high   concentrations   of   folding   intermediates   may   be   involved   in   the   process.(64,   65)     So   far   no   case   has   been   observed   where   both   the   monomer   and   dimer   of   a   wild-­‐type   protein   are   expressed   without   rapid,   equilibrating   monomer/dimer   exchange.   Domain-­‐ Swapping   can   also   be   caused   by   mutation.   For   example   a   single   site   mutation   in   protein   L   from   Peptostreptococcus  Magnus  leads  to  3D  domain  swapping  (66).     As   discussed   in   the   previous   chapter,   iLBPs   are   relatively   small   (126-­‐140   amino   acid,   predominantly  β  sheet  proteins)  cytosolic  proteins.  Structures  of  a  number  of  iLBPs  have  been   determined   and   all   share   this   monomeric   fold.     (Figure   II-­‐1)   Numerous   studies   of   the   folding   mechanism  of  members  of  the  iLBP  family  have  been  conducted,  specially  focused  on  human   Cellular   Retinoic   Acid   Binding   Protein   I   (hCRABPI).   (67-71)     In   an   elegant   series   of   experiments,   using   a   variety   of   techniques,   the   Gierasch   lab   has   concluded   that   CRABPI   undergoes   early   beta   barrel  collapse.   32   They   used   CRAPI   with   its   simple   architecture   to   study   structural   basis   of   β   sheet   proteins   formation.   Regarding   this   purpose,   three   only   Tryptophan   residues,   with   structurally   distinct   locations   as   three   probes,   were   mutated.   Folding   of   these   mutants   was   examined   using   stopped-­‐flow   fluorescence   and   circular   dichroism.   As   a   result   within   10   ms,   the   rapid   hydrophobic   collapse   occurs   and   stopped-­‐flow   circular   dichroism   shows   significant   secondary   structure   content   and   it   (chain   topology)   gets   developed   by   100   ms.   Then   followed   by   the   specific   packing   of   the   β-­‐sheet   sidechains   and     formation   of   the   native   hydrogen-­‐bond   network   (T  =   1   s).   These   are   hallmarks   of   the   pathway,   thus   preventing   off   pathway   aggregation   and   potential  amyloid  formation  in  this  protein.(72)   Investigations   into   the   folding   mechanism   of   other   members   of   the   family   have   also   been   conducted,   all   of   which   are   consistent   with   a   relatively   rugged   energy   landscape,   indicating  a  number  of  possible  stable  or  meta-­‐stable  intermediates,  a  situation  that  is  generally   characteristic  of  proteins  that  are  largely  beta  sheet  in  structure.(72)     Also   in   chapter   II   was   mentioned   that   CRBPs   belong   to   the   iLBP   family.   (6,   10-­‐13)   The   combination  of  small  size  and  relatively  large  binding  cavity  made  these  family  members  ideal   templates  for  use  in  protein  design  applications.   The   Geiger   and   Borhan   labs   have   used   these   templates   for   a   wide-­‐ranging   protein   design   applications   for   more   than   a   decade.   First   both   hCRABPII   and   hCRBPII   were   modified   to   be   Rhodopsin   mimics   that   bind   and   react   with   retinal   to   form   the   protonated   Schiff   base   (PSB)   which  is  found  in  the  binding  cavity  of  all  rhodopsins.(31,  73)  Then  the  environment  around  the   bound  chromophore  was  modified  to  study  the  mechanisms  of  absorption  wavelength  tuning   that   give   rise   to   color   vision,   and   the   light   selectivity   of   all   the   other   rhodopsins,   which   is   33   required  for  their  function.(74-­‐76)  Rhodopsin  mimics  that  photo-­‐isomerize  both  in  solution  and  in   the  crystal  were  created  to  allow  the  detailed  study  of  isomerization  processes  in  rhodopsins.(77,   78)   In   our   collaboration,   we   have   also   created   a   new   class   of   fluorescent   proteins   by   binding   fluorophores   in   the   binding   pocket   that   also   react   to   form   PSB’s   and   allow   regulation   of   the   emission  wavelength.   (79)  (Some  of  this  work  is  described  further  in  chapter  IV)  To  perform  all   the  above-­‐mentioned  projects,  we  had  to  make,  express  and  evaluate  hundreds  of  mutants  of   these  proteins.  During  the  preparation  of  mutants  for  these  studies,  Dr.  Wang  and  Dr.  Nossoni   for   the   fist   time   observed   protein   bands   that   eluted   separately   on   ion-­‐exchange   chromatography,   but   nonetheless,   were   the   same   species   on   SDS-­‐PAGE.     This   led   us   to   the   discovery  of  a  set  of  hCRBPII  protein  mutants  capable  of  domain  swapped  dimerization.  It  was  a   surprise   to   discover   that   some   hCRBPII   mutants   predominantly   form   dimers   via   domain   swapping  instead  of  the  monomeric  form.  (45)  (Figure  II-­‐3)           2 equiv         nomer"   rmediate   domain swapped dimer Figure  II-­‐3:  The  hCRBPII  DS  dimer.   We   were   the   first   group   to   report   domain   swapped   structures   of   hCRBPII. (45)   The   extent   of   domain   swapping   is   large   (almost   50%   of   the   protein   sequence)   and   occurs   during   routine   34   protein  bacterial  over-­‐expression  at  16-­‐30  °C.  HCRBPII  domain  swapped  dimerization  is  favored   by  high  expression  levels  in  E.  coli  and  high  concentrations  in  vitro.    Most  importantly  protein   stability  studies  show  the  domain  swapped  dimer  to  be  a  less  stable,  kinetically  trapped  species   (depending  on  the  mutant).  The  resultant  dimer  structure  represents  a  fold  not  identified  in  the   SCOP  or  CATH  protein  fold  databases.    Together,  our  data  supports  the  hypothesis  that  folding  of   hCRBPII  occurs  via  a  structurally  ordered  “open  monomer,”  with  the  relative  orientation  of  the   N-­‐   and   C-­‐   terminal   structural   units   determining   the   propensity   for   dimerization.2c   The   existence   of   this   dimer   for   hCRBPII   suggests   a   folding   pathway   distinct   from   that   characterized   for   hCRABPI,  indicating  that  evolutionary  related  members  of  the  same  structural  family  may  follow   different  folding  pathways.     Later   on,   we   determined   structures   of   holo   hCRBPII   domain   swapped   dimer   variants.   Together,   these   structures   show   an   extremely   large   and   reproducible   change   in   the   relative   orientation   of   the   two   domains   of   the   dimer   upon   ligand   binding,   inviting   the   possibility   that   iLBP  domain  swapped  dimers  could  be  allosterically-­‐regulated  forms  of  these  proteins,  at  least   in   some   cases.   Ideally,   the   domain   swapped   dimer   may   represent   a   new,   biologically   relevant   structural   fold   for   the   iLBP   protein   family,   though   it   remains   to   be   seen   if   they   form   in   vivo,   and   have  distinct  functionality  relative  to  their  parent,  monomeric  forms.     II-­‐2  Origin  of  dimerization  in  hCRBPII     As   it   was   mentioned   above,   we   developed   hCRBPII   as   rhodopsin   mimics   to   study   the   mechanism  of  wavelength  tuning.  To  this  purpose,  the  binding  site  was  redesigned  to  bind  all-­‐ trans  retinal  and  form  a  PSB  via  the  Q108K:K40L  (KL)  double  mutant  by  Dr.  Wang  previously  (80).       35       2.53 Å     4.23 Å         Figure  II-­‐4:  Trajectory  of  Tyr  60,   Glu  72  (both  shown  as  stick  representations,   colored  blue)  in  WT  hCRBPII  with  all-­‐trans-­‐retinal.   The  residues  situated  around  the  chromophore  were  then  selected  for  systematic  mutation  in   an  effort  to  understand  the  mechanism  of  wavelength  tuning.  Tyr60,  conserved  only  in  CRBPs,   is  located  in  beta-­‐strand  C  and  approximately  4Å  from  the  center  of  the  polyene  chain  of  all-­‐ trans-­‐retinal.  (Figure  II-­‐4)  (17, 80-82).     The   mutant   Q108K:K40L:Y60W   (KLY60W)   was   generated   by   Dr.   Wang,   expressed   and   purified   as   previously   described   in   the   hopes   of   producing   a   π-­‐π   interaction   with   the   chromophore   that   could   potentially   delocalize   the   positive   charge   of   the   PSB   along   the   chromophore,  leading  to  shifted  spectra.(80)  In  its  purification  process,  Source-­‐Q  ion  exchange   chromatography   (IEX)   unexpectedly   gave   two   KLY60W   hCRBPII-­‐containing   protein   peak   fractions,   one   eluting   at   80mM   NaCl,   where   hCRBPII   mutants   typically   elute,   and   a   second   eluting  at  160  mM  NaCl   (45)  (Figure   II-­‐5a).  The  UV-­‐Vis  spectra  of  the  all-­‐trans-­‐retinylidene  PSB   complex  with  protein  from  each  of  the  fractions  gave  distinct  spectra.  The  80  mM  and  160  mM   protein   fractions,   when   bound   to   retinal,   showed   Îťmax   of   496nm   and   514nm,   respectively   (Figure   II-­‐5b,   5c).     At   lower   expression   temperature   (25°C)   significantly   more   of   the   80   mM   36   fraction   was   produced   while   a   higher   expression   temperature   (30°C)   gave   similar   amounts   of   each  species  (Table  II-­‐1)       a     b   c     d                   e   f           g     h     i         Figure  II-­‐5:  a.  IEX  chromatogram  of  the  KLY60W,  monitored  at  280nm.  b.  The  UV-­‐ Vis   spectra   of   the   80mM   salt   elution   (red)   and   the   160mM   salt   elution   (blue)   of   KLY60W   incubated   with   all-­‐trans-­‐retinal.   c.   The   UV-­‐Vis   spectra   of   the   refolded   80mM   eluent   (low   salt,   blue)   after   denaturation   and   refolding   at   0.03   mM   concentration   and   the   native   160mM   (high   salt,   green)   and   native   80mM   elution   (red),  all  bound  to  all-­‐trans-­‐retinal.  d.  The  size  exclusion  chromatogram  (Superdex   200  16/75  column)  for  dimer  (the  160mM  elution)  KLY60W,  after  maintaining  the   protein   at   room   temperature   for   several   days,   showing   only   the   monomer   size   peak.  e.  Gel  filtration  chromatogram  of  monomer  (the  80mM  elution)  of  KLY60W,   showing   only   the   dimer   fractions.   f-­‐i.   Ion   exchange   chromatographs   of   WT   and   various  hCRBPII  mutants,  monitored  at  280  nm.  f.  Y60I  at  30°C.  g.  Y60L  at  30°C.  h.   Y60L  at  25°C.  i.  WT  hCRBPII  expressed  at  30  °C.   37   Size-­‐exclusion  chromatography  (SEC)  showed  that  the  80  mM  protein  eluted  as  a  15kD   monomer,   and   the   160   mM   fraction   eluted   as   a   30kD   dimer.   No   equilibrium   between   the   monomer   and   dimer   forms   was   observed,   even   after   incubation   for   10   days   at   room   temperature  (Figure   II-­‐5d   and   5e).  Each  species  has  been  tested  to  achieve  more  information   about   their   stability.   Their   melting   curves   showed   the   dimer   to   be   less   stable   than   the   monomer   (denaturing   at   69°C,   explained   in   II-­‐7-­‐4   section,   Figure   II-­‐6c   and   31),   however   the   monomer  was  stable  to  95°C  (Figure  II-­‐6f  and  II-­‐31).  The  160  mM  salt  fraction  was  denatured   with  8M  urea  and  refolded  at  low  concentration  (0.03  mM)  (Table   II-­‐1).  The  UV-­‐Vis  spectrum  of   the   retinal   PSB-­‐bound   refolded   monomer   was   identical   to   the   originally   isolated   monomer,   indicating  the  refolded  proteins  to  be  structurally  similar  (Figure  II-­‐5c).                           38     0 0 -20 -10 -30 -20 -40 -60 Wild-Type Dimer Tm = 52 ÂşC -80 -40 Y60L-Dimer Tm = 56 ÂşC -50   KLY60W-Dimer Tm = 69 ÂşC -50 -60 -70 -70 -140 20 30 40 50 60 70 80 90 100 -80 20 30 40 50 60 70 80 90 100 -80 20 -60 Temperature (ÂşC) d   e   -25 -50   -65 -70 -75 -80 -20 -35 -40 Y60L-Monomer -45 -55 20 40 60 80 100 Temperature (ÂşC) 120 -10 120 -30 -40 KLY60W- Monomer   -50 -60 -50 -85 -90 60 80 100 Temperature (ÂşC) -30 CD (mdeg) Wild-Type Monomer 40 f   CD (mdeg) -55 CD (mdeg) -40 -120 -60   -30 -100 Temperature (ÂşC)   c   -10 -20   C 20 CD (mdeg)   b   B CD (mdeg)   a  A CD (mdeg)   -70 20 40 60 80 100 120 30 40 50 60 70 80 90 100 110 Temperature (ÂşC) Temperature (ÂşC) Figure  II-­‐6:  Thermal  melting  curves:  a.  WT  hCRBPII  dimer  (θ  measured  at  220  nm),  Tm  =   52   °C;  b.  Y60L  dimer   (θ  measured   at  227   nm),  Tm   =  56  °C;  c.  KLY60W   dimer  (θ  measured   at   220   nm),   Tm   =   69   °C.   d.   WT   hCRBPII   monomer   (θ   measured   at   220   nm);   e.   Y60L   monomer  (θ  measured  at  215  nm);  f.  KLY60W  monomer  (θ  measured  at  220  nm).  Note   that   most   of   the   secondary   structure   is   preserved   in   both   the   WT   and   KLY60W   monomers,   even   at   the   highest   temperature   achieved   (CD   intensities   of-­‐65mdeg   and   -­‐ 50mdeg   respectively).   Thus   melting   curves   can   only   give   a   lower   limit   for   the   TM,   consistent   with   the   view   that   the   monomers   are   more   stable   than   the   dimers.   The   apparent  lack  of  a  Tm  for  the  monomers  might  suggest  the  presence  of  thermally  induced   folding   intermediates.   Clearly   the   unfolding   process   is   not   monolithic   as   that   observed   for   the   dimeric   species.   This   lends   further   support   to   the   suggestion   that   the   open   monomer,   which   is   presumably   obtained   from   melting   of   the   dimer,   has   a   different   folding  trajectory  (and   consequently   a   different   unfolding   trajectory)   as  compared  to   the   monomeric  species.                   II-­‐3  Dimer  formation  of  other  hCRBPII  variants   At   first,   we   hypothesized   mutation   of   the   residue   60   in   the   KLY60W   series   was   most   likely  responsible  for  the  formation  of  the  dimeric  species  since  many  hCRBPII  proteins  mutated   at  the  40  and  108  positions  have  been  purified  in  monomeric  form  in  our  lab  previously(80).  To   this  purpose,  we  made  several  single  mutants  at  this  position  (Table  II-­‐2).     39   Y60W  mutant  yielded  the  same  monomer/dimer  ratios  as  KLY60W,  confirming  the  importance   of   position   60.   (45)     In   fact   most   mutations   resulted   in   increased   dimer   formation,   including   Y60F,  Y60L  and  Y60I,  however  the  Y60H  mutant  produced  no  dimeric  protein.  (Table   II-­‐2   and   Figure   II-­‐5f-­‐h)   Together,   this   data   emphasizez   the   fact   that   position   60   mutations   can   significantly   favor   the   formation   of   dimeric   species.     Position   60   lies   near   the   N-­‐terminal   end   of   strand   4,   and   is   not   a   position   conserved   in   iLBP   family   members,   though   it   is   a   Tyr   in   most   retinol  binding  proteins.(81)       Table  II-­‐ 1:   The   Effect   of   expression   temperature   and  in   vitro   refolding   protein  concentration     on  the  dimer/monomer    (D/M)  ratio.                    Protein   D/M  (%)a   D/M  (%)a   In  vitro   at  25  °C   at  30  °C   refolding b WT-­‐hCRBPII   0/100   10/90    Q108K:K40L:                      Y60W   30/70   60/40                        Y60L   80/20   90/10     0.03  mM   0.13  mM   0.3  mM   0.03  mM   0.6  mM   0.03  mM   0.13  mM   0.4  mM   Percent  of    D/M  (%)c   0/100   0/100   40/60   0/100   30/70   20/80   40/60   70/30   a.  Mass  ratio  of  dimer  to  monomer;  b.  Starting  concentration;  c.  Mass  ratio,  after  in  vitro  refolding.       We   decided   to   perform   several   in   vitro   refolding   experiments   on   the   WT   and   several   of   the   mutants  (Table  II-­‐1  and  Figure  II-­‐7),  to  better  understand  the  nature  of  the  dimerization  in  this   protein.       40     a   b   c           d   e   f         Figure   II-­‐7:   Size   exclusion   chromatography   after   protein   denaturation   and   refolding,   all   monitored  at  280  nm.  a.  Y60L  mutant  refolded  at  0.03  mM  concentration.   b.  Y60L  refolded  at   0.13  mM  concentration.  c.  Y60L  refolded  at  0.4  mM  concentration.  d.    WT  hCRBPII  refolded  at   0.03   mM   concentration.   e.   WT   hCRBPII   refolded   at   0.13   mM   concentration.   f.   WT   hCRBPII   refolded  at  0.3  mM  concentration.     In   each   case,   dimer   formation   was   favored   by   higher   refolding   concentrations   while   monomer  formation  was  favored  at  lower  refolding  concentrations.    This  correlated  with  our  E.   coli-­‐expression  results,  where  expression  at  higher  temperatures,  which  leads  to  higher  levels   of  expression  and  therefore  higher  concentrations  of  pre-­‐folded  proteins,  (30  °C  versus  25  °C,   Table   II-­‐1   and   Figure   II-­‐7).  Melting  experiments  also  showed  the  dimer  to  be  less  stable  than   the  monomer  for  two  of  the  three  variants  (WT,  KLY60W,  Figure   II-­‐6a,   6c,   6d   and   6f),  indicating   the  dimer  to  be  the  kinetic  product  (the  melting  curve  for   the  Y60L  monomer  was  not  readily   interpretable,  Figure  II-­‐6b  and  6e).  (45)         41   ! Table  II-­‐2:    Monomer/dimer  ratio  of  various  residue  60  mutants.  During  bacterial  expression  at   30  °C     Mutations Dimer/Monomer mass ratio (%)   WT 10/90 (no dimer at RT) Y60L 90/10 S55W 20/80 S55W:Y60L 60/40   Y60I 80/20   Y60Q 20/80   Y60D 30/70   Y60T No dimer   Y60W 50/50   Q108K:K40L:Y60F 40/60   Q108K:K40L:Y60W 60/40   Δ56:WT 20/80   F57G:Y60L 70/30   T56P-Y60L 70/30   Δ56:Y60L 90/10   E72A 50/50   Q108K:K40L:Y60H No dimer             42   Overall,   these   experiments   are   consistent   with   the   dimer/monomer   ratio   being   kinetically   controlled,  with  higher  concentrations  favoring  dimer  formation.  (45)  In  the  meanwhile,  we  were   investigating   the   possibility   that   WT   hCRBPII   could   also   form   a   dimeric   species   by   overexpressing  hCRBPII  at  the  dimer  favoring  induction  temperature  of  30  °C,  resulting  in  both   monomeric  and  dimeric  protein  (Figure  II-­‐5i).  To  our  knowledge,  this  was  the  first  example  of  a   domain   swapped   system   where   both   monomer   and   dimer   of   the   wild-­‐type   species   are   expressed  without  monomer/dimer  exchange  (Figure  II-­‐8).  (45)         a   b           Figure   II-­‐8:   a.The   size   exclusion   chromatogram   for   dimer   of   WT   hCRBPII,   after   maintaining  the   protein  at  room  temperature  for  several  days,  showing  only  the  dimer   species.  b.  Gel  filtration  chromatogram  of  monomer  of  WT,  showing  only  the  monomer   protein.    Binding  assays  also  show  that  this  dimer  has  a  similar  affinity  (in  nM  range,  compared   to  monomer)  for  retinol  and  retinal  as  well  (Figure  II-­‐9,  method  explained  in  II-­‐3-­‐2  section).         43     a   b Kd = 5.64 nM c   Kd = 180 nM Kd = 4.36 nM d Kd = 110 nM                                             Figure   II-­‐9:     Ligand   binding,   monitored   by   Tryptophan   fluorescence   quenching   of     monomeric  and  dimeric  hCRBPII.  a.  Monomer  WT  hCRBPII  with  all-­‐trans-­‐retinol.  b.     Dimer   WT   hCRBPII   with   all-­‐trans-­‐retinol.   c.   Monomer   WT   hCRBPII   with   all-­‐trans-­‐   retinal.  d.  Dimer  WT  hCRBPII  with  all-­‐trans-­‐retinal.               II-­‐4  Structural  Studies   II-­‐4-­‐1  Structural  analysis  reveals  an  extensive  domain  swapped  dimer   We  were  able  to  crystallize  and  determine  several  dimer  structures  to  understand  the   nature  of  this  new  form  of  hCRBPII.  (45)  (Figure   II-­‐3)  In  this  attempt,  structures  of  dimer  mutants   (KLY60W,  Y60W  and  Y60L)  and  wild  type  (WT)  were  obtained  respectively  (Table  II-­‐7).  The  same   extensive,  domain  swapped  dimer  was  observed  for  all  four  of  these  structures  (Figures  II-­‐3,  10,   44   11,   12a   and   12b).  The  domain  swapping  is  extensive,  involving  residue  1-­‐56,  which  includes  3   beta  strands  and  two  helices  (Figure  II-­‐3,  12a,  12b).           aa Arg58 Trp60   Phe57 Ser55   Asn59   Thr56   bb Phe57 Trp60 Asn59   Ser55   Thr56 Arg58     c c         d Figure   II-­‐10:   Crystal  structure  of  dimer  KLY60W  mutant  and  its  monmer  components  a.   Chain   A   of   KLY60W   (where   only   the   Thr56   psi   angle   differs   from   the   closed   monomer   form)   with   the   electron   density,   contoured   at   1.0   σ,   of   the   hinge   loop   region   encompassing   amino   acids   55-­‐60.     Atoms   colored   by   type,   N,   blue,   O,   red,   C,   cyan.   b.   Chain   B,   showing   the   electron   density,   contoured   at   1.0   σ,   of   the   hinge   loop   region   encompassing  amino   acids  55-­‐60.  Atoms  colored  by  type,   N,  blue,  O,   red,  C,  cyan.   c.   The   complete   structure   of   the  KLY60W  dimer:  In  chain   A  (cyan)  Trp60   is  pointing   inside  the   binding  pocket,  while  in  chain  B  (pink)  Trp60  is  pointing  toward  the  solvent.     45   This  corresponds  to  an  83%  extent  of  swapping  as  defined  in  the  domain  swapping  3D   knowledge   database(http://caps.ncbs.res.in/3dswap/index.html).(83)   Each   of   the   two   domains   in   the   dimer   are   highly   similar   to   that   of   the   monomer   hCRBPII   structure   with   RMSD   ranging   from  0.382-­‐  0.397  between  monomeric  hCRBPII  and  a  single  domain  of  the  dimer  (Figure   II-­‐11).   These  domain  swapped  dimer  structures  have  a  large  rotation  about  the  psi  angle  of  Thr  56  in   common.                 Figure   II-­‐11:   An   overlay   of   the   dimeric   (cyan)   and   monomeric   (red)   KLY60W   mutant  structures.  Res56  for  both  structures  is  shown  in  sticks.  The  hinge  loop     region   in   the   monomer   and   dimer   is   encircled,   showing   the   dramatic   difference  in  the  two  structures.       While  the  psi  angle  of  Thr56  is  about  -­‐12°  in  monomeric  hCRBPII,  it  ranges  from  132°  –   169°  in  the  domain  swapped  dimers  (Figure   II-­‐12a)   This  150°+  rotation  about  the  psi  angle  of   Thr56   represents   the   hinge   motion   required   for   domain   swapping   (Figure   II-­‐11).   Though   mutation  at  residue  60  substantially  affects  the  monomer/dimer  ratio,  it  is  not  in  the  hinge  loop   region,   but   is   instead   3   amino   acids   away,   near   the   middle   of   β-­‐strand   4th.   In   fact   substantial   deviations   in   torsion   angles   from   the   monomer   structure   can   be   seen   in   residues   55   –   60   (from   the  hinge  loop  region  till  even  the  first  3  amino  acids  of  the  4th  beta  sheet).     46   a   b             c     d             e             Figure   II-­‐12:   a-­‐e:   Torsion   angle   differences   between   monomer   and   dimer   differences  of  Phi  angle  are  in  blue  and  psi  angle  differences  are  in  red  in  each     case.    a.  Thr56.  b.  Phe57.  c.  Arg58.  d.  Asn59.  e.  Res60.           The  domain  swapping  orients  the  opening  of  the  two  binding  sites  to  face  one  another,   creating   a   continuous   and   large   internal   cavity   that   stretches   from   one   side   of   the   dimer   to   the   other  (nearly  40  Å  in  length,  (please  refer  to  section  II-­‐6).  A  search  of  the  various  structural  fold   databases  (SCOP  1.75,  CATH  3.5)  reveals  no  other  protein  of  similar  architecture.(84)   47   II-­‐4-­‐2  Symmetry  VS  asymmetry     Although   all   four-­‐dimer   structures   shared   the   same   domain   swapped   interface,   there   are   substantial   differences   in   their   structures.   As   crystallographers   and   structural   biologists,   we   believed  that  the  structural  details  of  the  DS  dimers  are  key  to  understanding  the  mechanism  of   domain  swapping.     a     b         c d       Figure   II-­‐13:   The   differences   in   the   asymmetric   and   symmetric   structures   of   hCRBPII   dimers.  Res60   in   all   structures  is   shown   in   sticks  a.   The   asymmetric  structures:   The   overlay   structure   of   domain   swapped   dimer   KLY60W   (cyan)   and   Y60W   (pink).   In   chain   A   dimer   KLY60W,   Trp60   is   pointing   inside   the   binding   pocket.   b.   The   symmetric   structures.   The   overlay   of   WT   hCRBPII   (purple)   and   Y60L   (yellow)   domain   swapped   dimer.   c.   Comparing   asymmetric  (dimer  KLY60W  in  cyan)  and  symmetric  dimer  (dimer  Y60L  in  yellow)  by  looking   down  the  two-­‐fold   axis   and   it   shows  the   different   position   of   helices   in   N-­‐terminus.   d.   One   chain  of  dimer  Y60L  (yellow)  and  both  chains  of  asymmetric  dimer  Y60W  (chain  A  in  cyan   and  chain  B  in  red)  are  overlaid  at  their  C-­‐terminus.   We   found   that   hCRBPII   DS   dimers   can   be   classified   as   either   symmetric   (WT   hCRBPII,   Y60L   and   other   mutants   described   in   II-­‐6   section),   where   the   two   chains   which   make   the   dimer   are  essentially  identical,  or  asymmetric  (KLY60W,  Y60W  and  holo  DS  dimers  described  in  II-­‐6),   where  the  two  chains  are  distinct  from  one  another  (Figure   II-­‐13). (45)    As  shown  in  Figure   II-­‐13b   48   the  two  symmetric  structures  (WT  and  Y60L)  are  similar,  with  the  relative  orientation  of  the  two   domains   in   the   dimer   essentially   identical.     However,   there   are   significant   differences   in   the   torsion  angles  around  residue  60,  with  an  associated  repositioning  of  the  side  chain.    Tyr60  is   buried  inside  of  the  protein,  making  a  hydrogen  bond  with  Glu72  in  both  the  WT  monomer  and   dimer,  while  Leu60  is  flipped  out  of  the  inside  of  the  protein  and  solvent  exposed  in  the  Y60L   dimer  (Figure  II-­‐14a).     This   “flipped   out”   conformation   is   the   cause   of   the   residue   59   phi   angle   large   deviation,   relative   to   its   monomer   value   (about   -­‐121°   in   the   monomer   versus   -­‐80°   in   the   Y60L   dimer)   (Figure  II-­‐14c)  In  contrast,  the  phi  torsion  angle  of  Asn59  in  the  WT  dimer  is  -­‐142°  (similar  to  its   monomer   value),   which   leads   to   the   “flipped   in”   conformation   (Figures   II-­‐14a   and   14b).     To   balance   the   resulting   change   in   trajectory   of   the   main   chain,   the   Asn59   psi   angle   is   radically   rotated   relative   to   all   of   the   other   structures   (-­‐142°   in   the   WT   dimer   versus   161°   in   the   monomer  and  165  in  the  dimer  Y60L,  see  Figures  II-­‐b),  resulting  in  a  substantial  Ramachandran   outlier  at  Asn59.  Nonetheless  it  results  in  the  main  chains  of  the  two  symmetric  structures  once   again  closely  tracking  each  other  (Figure  II-­‐14a).       49   a     2.6 Å   Tyr60 Glu72 Leu60   Asn59   b   c     Glu72 Glu72 2.6 Å Tyr60       Leu60 Asn59 ÎŚ = -142Âş Ψ = -142Âş (161Âş in monomer) Asn59 ÎŚ = -80Âş (-121Âş in monomer) Ψ = 165Âş Figure   II-­‐14.   Torsion   angle   deviations   outside   the   hinge-­‐loop   region   define   the     relative  orientation  of  the  two  domains  of  the  dimer,  with  the  highest  deviation   seen  in  Asn59.  a.    The  WT  (purple)  and  Y60L  (yellow)  dimers  are  overlaid.    Inset,     the   hydrogen   bond   between   Tyr60   and   Glu72   compared   to   the   “flipped   out”   conformation   of  Leu60.  b.  The   critical  Tyr60  and  Asn59  region   in  the  WT   hCRBPII     dimer,   showing   the   key   phi/psi   angles.   c.   The   same   region   in   the   Y60L   dimer,   showing  the  “flipped  out”  Leu60  and  the  key  phi/psi  angles.    Comparison  of  the     two  shows  how  the  large  difference  in  the  phi  angle  is  compensated  for  in  the  WT   N59  psi  angle,  keeping  the  main  chain  of  the  two  on  a  similar  trajectory.       In   contrast   to   the   symmetric   dimers,   asymmetric   dimer   (i.e.   Y60W   and   KLY60W)   structures   are   similar   to   one   another   and   asymmetric   (Figures   II-­‐13a).   However,   there   is   a   significant  deviation  in  the  position  of  the  N-­‐terminal  region  when  the  C-­‐terminal  domains  are   overlaid  (Figures   II-­‐13c   and   13d).  Basically,  in  asymmetric  dimers,  there  is  a  large  deviation  in   the   relative   orientation   of   the   N-­‐   and   C-­‐terminal   regions   in   the   two-­‐polypeptide   chains   (subunits)  that  make  the  dimer.  (Figure  II-­‐13d) (45)   50     V62   R58 W60 F57   D63 N59   D61 T56   V62   N59 W60 F57   D63 D61   R58 T56   Figure   II-­‐15:   Residues   56-­‐63   in   the   chains   A   (green,   top)   and   B   (magenta,     bottom)  of  Y60W  hCRBPII.       This  positional  deviation  is  as  large  as  14  Å  between  the  two  conformations  of  the  DS   dimer.    This  represents  a  radical  change  in  the  relative  orientation  of  the  two  domains  relative   to  the  symmetric  dimer.       The   source   of   this   large   deviation   can   be   seen   in   the   torsion   angles   of   the   residues   between  56-­‐60  (Figures  II-­‐12).    Where  in  chain  A  of  both  KLY60W  and  Y60W  dimers,  all  of  the   torsion  angles,  except  the  psi  angle  of  Thr56,  are  similar  to  those  of  a  monomer.  In  chain  B  a   number  of  other  torsion  angles,  namely  the  psi  angles  of  residues  Thr56  and  Phe57,  and  the  phi   angle   of   Phe57,   deviate   from   their   monomer   values   (Figures   II-­‐12a   and   12b).     Presumably,   these  differences  in  torsion  angles  are  required  to  orient  the  two  halves  of  the  chain  properly   for  dimer  formation.   51   Interestingly,  the  A  subunit  of  the  KLY60W  and  Y60W  mutants  is  the  most  similar  to  the   monomer.  It  has  phi/psi  angles  almost  identical  to  those  of  the  monomer,  except  the  psi  angle   of  hinge  residue  T56,  which  is  rotated  by  almost  180  degrees  in  all  the  DS  dimers.  Phe57  and   Arg58  side  chains  are  on  the  same  side  of  the  strand,  similar  to  the  conformation  seen  in  the   loop  connecting  the  two  strands  in  the  monomer   (Figure   II-­‐12c   and   15).   However,  the  phi/psi   angles  of  the  B  subunit  had  other  significant  differences  with  monomer,  especially  the  phi  angle   of   residue   59   (Figure   II-­‐12d),   which   resulted   in   the   flipping   of   Trp60   from   inside   the   binding   cavity  to  the  outside  and  Asn59  from  outside  to  inside  of  the  cavity  (relative  to  the  monomer)   (Figure   II-­‐15).   This  places  both  residues  60  and  61  outside  the  binding  cavity,  which  is  necessary   to   re-­‐phase   the   strand   (explained   in   the   next   section,   II-­‐4-­‐2).   The   asymmetry   of   the   Y60W   mutant  dimers  tells  us  that  the  “canonical”  DS  dimer,  where  the  only  torsion  angle  change  is  in   the   hinge   residue,   is   not   possible   because   the   relative   orientations   of   the   N-­‐   and   C-­‐terminal   regions   of   each   domains   are   incompatible   with   dimer   formation;   thus   subunit   B   must   adjust   itself,   via   the   torsion   angles   along   the   connecting   strand,   to   adopt   a   conformation   consistent   with  dimer  formation  to  subunit  A.(45)  (Figure  II-­‐16) 52         Figure   II-­‐16:   Overview  of  dimer  formation  in  asymmetric  structures:  Two  chains  of   Y60W  dimer  are  labeled  in  the  picture.  The  same  chain  (molecule  A  and/or  B)  in   this   asymmetric   dimers   cannot   form   dimers   since   they   clash.   However,   the   interaction   of   different   chain   with   each   other   (molecule   A   and   B)   could   lead   to   proper  DS  dimer  formation  with  the  different  orientation.   II-­‐4-­‐3  Phase  Relationship     In  all  cases  domain  swapping  is  the  result  of  the  straightened  connecting  loop  between   beta   strands   3     and   4   of   the   monomer,   resulting   in   this   loop   and   beta   strands   3rd   and   4th   becoming  a  single  beta  strand  stretching  the  length  of  the  dimer  (Figure  II-­‐3).         53     N59 D63 D61 = Carbonyl C =ÎąC = Nitrogen V62 Y60 R58 = β Sheet, Side Chain In T51 T53 S55 K50 K52 T54   F57 T56 = β Sheet, Side Chain Out = Loop   Monomer   T51   K50 T53 K52 S55 T54 F57 T56 V62 N59 R58 Res 60 D61 D63 Dimer   Figure   II-­‐17:   Schematic   representation   of   “phase   relationship”   in   DS   dimerization.   Top:  monomer,  bottom:  dimer  with  residue  60  as  mutant.     In  order  to  have  a  DS  dimer,  with  a  perfect  stretched  beta  sheet  in  that  region,  eventually  we   will  have  “phase  problem”  because,  in  monomer  while  the  odd  side  chains  of  beta  strand  3  face   inside   the   binding   pocket,   the   even   side   chains   of   beta   strand   4   face   inside.   With   an   even   number  of  residues  in  the  loop,  this  means  that  the  single  beta  strand  formed  in  the  DS  dimer   must,  at  some  point,  re-­‐phase  the  strand,  which  would  then  put  the  even  side  chains  of  the  C-­‐ terminal  half  of  the  strand  “in  phase”  with  the  odd  numbered  side  chains  of  the  first  half.     The  way  that  mutants  (i.e.  Y60L  and  Y60W)  solve  the  phase  problem  might  be  the  key   point   in   the   domain   swapping   mechanism.   Conformations   of   residues   Y60   and   D61   are   important   to   consider   as   well.   In   Y60L,   Y60W   (chain   B),   KLY60W   (chain   B)   mutant   domain   swapped  dimers  of  hCRBPII,  these  two  residues  are  pointed  toward  the  solvent,  which  re-­‐phase   the  strand  in  order  to  fix  the  phase  problem  (figure   II-­‐17).  However,  in  wild  type  hCRBPII,  which   gave  us  mostly  monomer,  residue  Y60  is  inside  the  binding  pocket.  This  may  suggest  one  of  the   54   significant  points  in  the  mechanism  for  dimerization  in  hCRBPII. II-­‐5  possible  mechanism  for  domain  swapping  in  hCRBPII     With   the   existence   of   the   WT   domain   swapped   dimer   and   the   fact   that   there   is   no   Interconversion  between  monomer  and  dimer,  indicating  each  as  a  unique  fold  of  this  protein.       Therefore,  we  proposed  a  mechanism  for  the  hCRBPII  folding  pathway,  different  from  what  has   been   seen   for   other   iLBPs   earlier(72).     We   suggested   that   the   two   halves   of   hCRBPII   (the   N-­‐ terminal  and  C-­‐terminal  halves)  are  capable  of  at  least  partially  folding  independently,  initially   as   an   extended   “open   monomer.”     The   dimer/monomer   ratio   would   then   depend   on   the   relative   rates   of   dimerization   of   the   open   monomers,   versus   rotation   of   the   n-­‐   and   c-­‐termini   together  to  form  the  “closed  monomer”.   Our  data  shows  that  three  things  are  required  for  domain  swapping  in  this  system,  the   large  hinge  motion  at  Thr56,  subsequent  proper  orientation  of  the  n-­‐  and  c-­‐termini  to  adopt  a   conformation  consistent  with  dimer  formation  and  re-­‐phasing  of  the  connecting  strand  with  the   last  two  strongly  correlated.  (Figure  II-­‐18)             55   T56 F57 HOOC   NH2     H2N COOH Folded Monomer   NH2   NH2 H2N COOH Domain Swapped Dimer   T56   T56 F57 F57           HOOC Open Monomer COOH Rephased Open Monomer Figure   II-­‐18:   DS   dimerization   requires:   1.   Rotation   about   Thr56   psi.   2.   Orientation   of   the   two   halves   to   accommodate   dimerization.     3.   Rephasing  of  the  connecting  strand.   As  one  would  expect,  this  is  a  rate-­‐governed  process,  thus  factors  such  as  temperature   and  concentration  should  affect  the  outcome.  In  all  the  symmetric  dimers  thus  far  structurally   characterized   virtually   the   same   dimer   is   seen,   where   both   subunits   of   the   dimer   have   very   similar   relative   orientations   of   their   n-­‐and   c-­‐termini,   with   Asn59   flipped   inside   the   binding   cavity,  which  is  “in  phase”  with  the  n-­‐terminal  side  of  the  connecting  beta  strand.  (Figure  II-­‐21)   However,   reaching   this   symmetric   dimer   arrangement   is   accomplished   in   different   ways.   Flipping  out  of  the  residue  at  position  60,  in  both  the  Y60L  symmetric  dimer  and  molecule  B  of   both   asymmetric   dimers   (Figures   II-­‐14a   and   15)   is   one   low   energy   pathway   for   a   “dimer   friendly”  orientation.  In  contrast  WT  hCRBPII  is  “spring  loaded”  against  this  conformation  by  the   56   hydrogen  bond  made  between  Glu72  and  Tyr60,  which  serves  to  hold  Tyr60  in  the  “flipped  in”   position  (Figures  II-­‐14b).       Formation   of   the   WT   symmetric   dimer   then   requires   an   almost   90°   rotation   of   the   Asn59  psi  angle  resulting  in  a  substantial  Ramachandran  outlier  (Figure  II-­‐12d).  (“Twist  against   the   spring”   situation).     This   gave   rise   to   the   small   quantities   of   WT   dimer   and   its   relative   instability.  Tyr  at  position  60  is  resistant  to  flipping  out  due  to  the  hydrogen  bond  with  Glu72,   which   “spring   loaded”   it   against   dimerization,   (Figure   II-­‐19)   and   therefore   requiring   an   unfavorable   psi   angle   in   Asn59   to   achieve   the   relative   orientation   required   for   dimerization   (Figure  II-­‐  12d).                         Figure   II-­‐19:   DS   dimer   vs.   monomer   formation   in   WT   hCRBP  II.     In   none   of   the   other   DS   dimer   structures   we   see   Ramachandran   outliers   in   the   connecting   strand   region,   indicating   since   Tyr60   is   inside   the   binding   pocket   and   making   its   hydrogen   bond   with   Glu72   results   in   a   relatively   high   energetic   penalty   for   dimerization,   due   57   both  subunits  to  the  need  to  rephase  the  sidechains  on  the  connecting  strand,  and  to  allow  the   n-­‐  and  c-­‐termini  to  adopt  a  conformation  consistent  with  dimerization.     Based   on   this   hypothesis   mutation   of   E72   to   a   residue   incompetent   for   hydrogen   bonding   (a   small   hydrophobic   residue)   will   “loosen   the   spring”   resulting   in   an   increase   in   dimer   formation.  (Figure  II-­‐19)       II-­‐5-­‐1  Dimerization  and  structure  of  the  E72A  mutant   As   a   first   step   to   confirm   and   test   the   abovementioned   hypothesis,   the   E72A   mutant   was   created,   expressed   and   purified.     As   predicted   this   mutation   led   to   a   dramatic   increase   in   dimerization   relative   to   WT   hCRBPII.     (E72A   mutant   produced   50%   dimer   vs.   only10%   in   WT)   (Table  II-­‐9,  Figure  II-­‐20).    The  E72A  mutant  was  then  crystallized  and  its  structure  determined.       a   b           Figure   II-­‐20:  Size  exclusion   chromatography  of  low  salt  and   high   salt   (80   mM   and   160   mM   NaCl,)  E72A   hCRBPII   respectively,   expressed   at  25°C  (since  30°C  expressed   produced  virtually  no  soluble  protein).     Although  a  crystallographic  two-­‐fold  axis  does  not  relate  the  two  monomers  in   this  dimer,  the  relative  orientation  of  the  N-­‐  and  C-­‐terminal  halves  of  the  protein  are  essentially   identical   to   that   of   the   symmetric   dimers   (Figure   II-­‐21C).   There   are   deviations   in   its   torsion   58   angles  at  position  57  (phi  angles  of  -­‐85°  versus  -­‐133)  and  most  notably,  position  60  (phi  angles   of   35°   versus   -­‐72°   and   psi   angles   of   92°   versus   127)   (Figure  II-­‐12b  and  e).     The   result   is   that   Tyr   60   is   in   the   “flipped   out”   conformation,   similar   to   that   seen   in   both   the   Y60L   dimer   and   no   Ramachandran   outliers   in   the   connecting   strand   residues   (Figure   II-­‐21a),   thus   inducing   the   flexibility   between   the   n-­‐   and   c-­‐terminal   domains   (compare   to   WT,   Figure   II-­‐21b),   favoring   dimer  formation,  and  suggesting  that  the  hydrogen  bond  between  Tyr60  and  Glu72  is  there  to   repress  dimer  formation  and  consistent  with  the  hypothesis  that  the  interaction  between  Y60   and   E72   in   the   open   monomer   folding   intermediate   effectively   holds   the   two   domains   in   a   conformation   unfavorable   to   dimerization,   promoting   the   formation   of   the   physiologically   relevant  monomeric  species.    This  suggests  that  the  n-­‐  and  c-­‐terminal  domains  are  capable  of  at   least  partially  folding  in  isolation,  and  would  represent  the  “foldons”(85-­‐89)  for  hCRBPII.  (Figure  II-­‐ 18  and  19)  Note  that  Tyr60  is  not  conserved  in  the  iLBP  family,  and  is  found  only  in  CRBP’s.                 59   a   a   Glu72 Ala72   Tyr60 Leu60 Asn59     b b   Ala72 Tyr60   Asn59 Tyr60 Glu72 2.6 Å     c c         Figure   II-­‐21:   a.   Overlay   of   chain   A   of   the   E72A   dimer   (green)   and   the   Y60L   dimer     (yellow).  Note  that  both  show  residue  60  pointing  toward  the  solvent.  b.  Overlay  of   the  E72A  chain   A  (blue)  and  WT  (green)   dimers.   Inset  shows  that  without   the  Glu  72     hydrogen   bond  donor,  residue  60  is   free  to  flip  out,  leading   to   a  more  relaxed  dimer   structure.   c.   Two   chains   of   the   E72A   dimer   (green   and   grey)   are   overlaid,   showing     their  strong  similarity.     II-­‐5-­‐2  Study  the  folding  pathway     As   mentioned   above,   we   proposed   that   the   N   and   C   terminus   of   the   protein   might   fold   independently     (at   least   partially)   to   form   an   open   monomer   folding   intermediate.   Then,   the   ratio   of   monomer/dimer   would   depend   on   the   relative   rates   of   dimerization   of   open   monomers,  versus  rotation  of  the  N-­‐  and  C-­‐termini  together  to  form  the  “closed  monomer”.  It   60   was   seen   before   that   this   is   concentration   dependent   (Table   II-­‐1)   and   would   depend   on   the   energetics  of  proper  orientation  for  dimer  formation.     In  this  Scenario,  the  DS  dimer  would  then  represent  the  kineticly  trapped  folding  intermediate   on  the  native  folding  pathway  of  hCRBPII. One  of  the  early  steps  to  detect  and  study  the  folding   intermediates   is   to   monitor   the   protein   at   different   denaturant   concentrations,   with   the   assumption   that   the   structures   formed   at   intermediate   denaturant   concentrations   will   reflect   the   intermediates   formed   in   the   folding   pathway.(90,   91)   To   this   end,   we   tried   to   probe   the   structure  of  WT-­‐hCRBPII  over  a  range  of  denaturant  concentrations  (Gd-­‐HCl  in  this  case)  using   both   Circular   dichroism   spectroscopy   and   tryptophan   fluorescence   spectroscopy.(1)   (Figure   II-­‐ 22)   We   initially   were   interested   to   plot   the   change   in   ellipticity   versus   denaturant,   (Figure   II-­‐ 22c)  to  compare  it  to  a  two-­‐state  (N  ⇄  U)  and  three  state  (N  ⇄  (1/n)In  ⇄  U  or  even  higher  order   model)  which  would  suggest  the  presence  of  intermediates  (excluded  the  two-­‐state  model)  in   the   folding   pathway(92).   It   should   be   also   considered   that   the   two   halves   of   the   protein   may   have   significantly   different   stabilities.   After   performing   these   series   of   experiments,   the   helix/strand   ratio   was   calculated   from   the   CD   spectra   at   each   individual   denaturant   concentration   and   compared   to   the   secondary   structure   of   the   native   and   predicted   intermediate  (which  almost  consists  of  the  same  percentages  of  helix/strands  as  native)  (Figure   II-­‐19  and  Table  II-­‐8).(1)  Also  the  concentration  dependence  may  suggest  that  dimer  association   occurs   early   in   the   protein-­‐folding   pathway.     Based   on   the   result   from   our   CD   spectra   and   calculated  percentages  of  helix  and  beta  sheet  at  different  Gd.Hcl  concentrations,  helixes  may   be  more  stable  compare  to  beta  sheets.  From  1M  of  denaturant  concentration,  the  percentages   of  beta  sheets  decrease  to  30%  while  for  helixes,  even  at  2M  still  there  are  intensities  relative   61   to  ι-­‐helix  signal  and  after  4M  it  drops  to  about  20%  of  the  native  (Table  II-­‐8),  proposing  that  N-­‐ terminus   maybe   (partially)   fold   first.   These   calculations   are   all   considered   with   the   respect   of   220nm  regarding  beta  sheets  and  222nm/208nm  for  helixes.           a c             b d         Figure   II-­‐22:   Equilibrium   Unfolding   Experiment   for   wild-­‐type   CRBPII   in   presence   of   different   concentration   of   Gd-­‐HCl,   a.-­‐b.   monitoring   by   Circular   Dichroism   (CD).   c-­‐d.   Tryptophan  fluorescence  spectroscopy,  right,  performed  with  WT  hCRBPII,  to  study  the   presence   an   intermediate   in   the   folding   pathway.     Purple   solid   line   represents   the   polynomial   fit  (3rd  degree)  in  b  and  d.  Both  b   and   d  spectra   are  in  correlate  with   previous   studies.  (1)             62   Also,   based   on   the   previous   studies   on   analysis   of   three-­‐state   protein   unfolding   data,   our   CD   spectra   of   unfolding   supports   more   three-­‐state   model   than   two   state.   (1)     Even,   Prof.   Lapidus   have   tried   to   fit   our   data   into   two-­‐state   model   and   it   did   not   fit   correctly.     However   further   analysis,  especially  with  three-­‐state  model  fit,  need  to  be  done  in  order  to  test  this  hypothesis.   Same  set  of  experiments  can  be  done  with  the  dimer  favoring  mutants  (i.e  Y60L),  and  conduct   them   at   various   monomer-­‐favoring   and   dimer-­‐favoring   concentrations.   Observing   a   similar   intermediate  in  the  folding  of  these  mutants  confirms  our  mechanism  even  though  the  folding   product   is   different   and   it   would   allow   us   manipulating   the   folding   pathway   to   control   the   folding  product  for  this  protein,  and  potentially  other  members  of  the  iLBP  family  as  well.             Completing   mentioned   calculations   requires   more   recourses   period,   however   during   the   course  of  my  graduate  research,  were  unable  to  complete  the  mentioned  task.  Though  we  are   optimistic  this  project  will  carry  on   II-­‐5-­‐3  HCRBPII  folding  route  VS  other  iLBP  members   A   number   of   folding   studies   have   been   conducted   on   a   few   other   iLBP   family   members,   namely   hCRBPI   and   human   cellular   retinoic   acid   binding   protein   I   (hCRABPI).     In   the   latter   case,   a  series  of  experiments  have  indicated  that  the  folding  pathway  of  hCRABPI  involves  an  early   barrel   closure   in   the   folding   pathway,   even   before   substantial   secondary   structure   is   evident(67-­‐ 71) .             63                   Figure  II-­‐23:  overview  picture  of  hCRABPI  folding  pathway  compare  to  hCRBPII.     This  prevented  the  protein  from  forming  intermediates  that  were  prone  to  aggregation   and  potentially  fibril  formation.  This  mechanism  is  clearly  distinct  from  what  we  proposed  for   hCRBPII.  However,  there  is  no  guarantee  that  all  members  of  a  structural  family  would  follow   the   same   folding   pathway.   Recently,   the   crystal   structure   of   a   FABP5   DS   dimer   was   reported.     The  swapped  region  was  almost  identical  to  that  of  hCRBPII.(93)  (Refer  to  next  Chapter,  IV)  This   shows   that   other   family   members   ((with   only   36%   sequence   identity   between   FABP5   and   hCRBPII)  may  also  have  a  propensity  to  form  the  same  DS  dimer.  The  identification  of  a  second,   DS   dimer   iLBP   family   member   lead   us   to   the   idea   that   a   subset   of   iLBP   family   members   may   indeed  be  DS  dimers  in  their  physiologically  relevant  forms.   II-­‐6  Conformational  change  driven  by  ligand  binding  in  hCRBPII   It  was  previously  mentioned  that  the  cellular  retinol  binding  proteins  (CRBP’s)  are  involved  in   the  trafficking  of  both  retinol  and  retinal  within  the  cell.  Binding  studies  of  hCRBPII  DS  dimers   64   have  demonstrated  that  DS  dimers  are  competent  for  ligand  binding.(45)  (Figure   II-­‐9)  Knowing   these,   in   continues   collaboration   between   our   lab   (prof.   Geiger   group   members)   and   Prof.   Borhan   group;   we   successfully   expressed,   purified   and   determined   structures   of   multiple   variants  of  holo  hCRBPII  domain  swapped  dimer  mutants  (up  to  90%  dimer  formation)  bound   Protonated  Schiff  Base,  PSB,  to  retinal  as  one  of  its  natural  ligand  and  in  the  absence  of  ligand  in   that   series.   (74-­‐76)c     (Figure   II-­‐   25,   II-­‐28   and   II-­‐29)   Q108K:T51D   holo   structure   obtained   by   Dr.   Nosrati  and  apo  form  obtained  by  Alireza  Ghanbarpour.  (Table  II-­‐  3)     We  found  that  ligand  binding  in  dimers  leads  to  a  reproducible  and  noticeable  conformational   change.   Among   those,   all   of   the   apo   structures   reveal   a   symmetric   dimer   very   similar   to   the   other   symmetric   dimers   previously   mentioned   in   section   II-­‐4-­‐2   (Figure   II-­‐24),   though   some   of   them  might  differ  in  their  crystal  packing.  Therefore,  the  data  indicate  the  symmetric  dimer  to   be   the   most   common   and   robust   form   of   DS   dimer   structure,   with   the   asymmetric   dimer   a   special  case  caused  by  the  Y60W  mutation.                                 65   Table  II-­‐3:    Summary  of  DS  dimer  crystal  structures  of  all  hCRBPII  mutants         #  molecules       Mutants   Holo/Apo   per   Symmetric/     Q108K:T51D   Q108K:T51D   Q108K:K40D   Apo   Holo   Holo   asymmetric   unit   asymmetric   1   12   12   Symmetric   Asymmetric   Asymmetric         On   the   other   hand,   structures   of   the   retinylidene-­‐bound   dimers   reveal   different   relative   orientation   of   the   two   domains   of   the   dimer   in   them   when   compared   to   the   apo   symmetric   dimers  (Figure  II-­‐25).                       Figure   II-­‐24:   A   comparison   of   four   symmetric   dimers,   Y60L   (orange),   WT   (red),  E72A  (cyan)  and  Q108K,K40D  (green).   Further,   most   of   the   holo   dimers   crystallize   with   between   4   and   12   monomers   in   the   asymmetric   unit,   all   of   which   exhibit   small   differences   in   relative   orientation   of   the   two   domains  that  make  up  the  DS  dimer.  This  indicates  that,  in  contrast  to  the  apo  DS  dimer,  which   seems   to   exist   as   a   relatively   rigid   structure,   holo   dimers   have   more   flexibility   along   their   66   strand.  However,  nearly  all  dimer  molecules  of  holo  structures  (i.e.  6  pairs  in  Q108K:K40D)  are   pretty  similar  to  one  another.  (Figure  II-­‐28a-­‐f)               11A#         Figure  II-­‐25:  Overlay  of  holo  Q108K:K40D  (chains  F  and  A,  both  shown    in  orange)  and  apo   Q108K:T51D   (shown   in   blue)   showing   the   large   motion   of   helix   1   upon   ligand   binding   (about   15  Å).  Bound  retinal  molecules  in  holo  structures  are  shown  as  transparent  spherical  models.   Note,   complete   dimer   in   the   apo   structures   was   generated   by   crystallographic   two-­‐fold   symmetry  operation.       The  major  cause  of  this  large  relative  deviation  in  two  domains  is  related  to  the  position   Thr59  and  Tyr60.    In  all  of  the  mentioned  apo  symmetric  dimers,  the  Asn59  side  chain  is  flipped   into  the  binding  cavity  and  the  residue  60  is  flipped  out  of  the  binding  cavity  (except  the  WT-­‐ hCRBPII  dimer  has  Y60  flipped  in  form).(94)  This  is  in  contrast  to  every  hCRBPII  monomer  (with   over   40   structures   so   far   determined),   where   Asn59   is   always   outside   the   binding   pocket.   In   holo  DS  dimer  structures  the  steric  bulk  of  the  ligand  forces  Asn59  to  rotate  out  of  the  binding   pocket,   resulting   in   the   rotation   of   the   entire   domain. (Figure  II-­‐26a  and  b),   and   Tyr60   is   found   inside  the  binding  pocket.     67     a     LYS108 b   Re  37,but  most  preferably  those  with  degree  of  polymerization  DP 6-15  to  link  multiple  clusters   of   amylopectin   5f,   while   BEIIb   specifically   transfers   short   chains   with   mostly   DP   of   6   and   7.   (Figure   V-­‐1)   In   contrast,   the   BEIIa-­‐deficient   mutant   exhibited   no   significant   change   in   the   amylopectin  chain.  (152)  The  physiological  role  of  BEIIa  might  be  to  support  at  least  partially  the   function  of  BEI  and  BEIIb.(151),  5f     The   structure   of   truncated   Oryza   Sativa   L   (Asian   rice)   branching   enzyme   (RBEI)   by   158   Noghuchi   et   al.   revealed   the   catalytic   residues   and   a   few   glucan   surface   binding   sites   (where   the   protein   interacts   and   binds   with   amylose)   were   observed   to   play   an   important   role   on   how   BEI  recognizes  polysaccharides  containing  ι-­‐1,4  as  well  as  ι-­‐1,6  linkages.  (2)                                 Figure  V-­‐1:  Schematic  view  of  the  BE  isozymes  and  SS  isozymes  in  the  amylopectin  cluster.     V-­‐2  Structure  of  the  both  N  and  C  terminus  truncated  RBEI:   The   crystal   structure   of   the   both   N   and   C   terminus   truncated   RBEI   first   has   been   determined   by   Noghuchi   et   al.   in   2011.   Also,   Dr.   Remie   Fawaz   (a   former   Geiger   lab   member)   159   has   obtained   another   truncated   structure   of   this   enzyme   recently   with   dodecaose   (M12).   (FigureV-­‐2)     BEI  is  820-­‐residues.  The  N-­‐terminal  leader  sequence  (65  residues),  which  is  responsible   for   transport   into   the   amyloplast,   is   truncated   in   the   structure   and   leads   to   mature   BEI.   A   mature   BEI   consists   of   755   residues   and   four   domains:   a   central   (β/Îą)8   catalytic   domain   (the   GH13   module,   residues   161–   587),   the   N-­‐terminal   carbohydrate-­‐binding   module   48   domain   (CBM48;  residues  59–160),  the  N-­‐terminal  helices  (residues  1-­‐58)  and  an  ι-­‐amylase  C-­‐  domain   (residues   588–702)   (FigureI-­‐2).   Asp344   and   Glu399,   the   essential   catalytic   residues,   are   the   nucleophile  (base)  and  proton  donor  respectively.  (2,  153)       Crystal   structures   of   both   rice   mature   BEI   and   BEIΔC   (residues   702-­‐755   are   truncated)   were  obtained  before.  (118)  In  our  group    Dr.  Fawaz  was  able  to  obtain  well  diffracting  crystals  of   malto-­‐dodecaose  (M12)    bound  BEIΔC  at  2.35  Å.         160   Active  Site   Binding  Site  I   Binding  Site  II   Figure   V-­‐2:   Overall   structure   of   rice   Branching   Enzyme   I   in   complex   with   maltododecaose   (M12).   The   N-­‐terminal   domain   is   shown   in   green,   CBM48,   pink,   center   catalytic   domain,   cyan,   and   the   C-­‐terminal   Îą-­‐amylase   C   domain   in   blue.   Carbohydrates   are   represented   in   sticks   and   colored   by  atom   type,   carbons   in   yellow   and   oxygens   in   red.   One   oligosaccharide,   M12,  binds  exclusively  into  the   catalytic  domain  and  hangs  over  the   catalytic  groove  without   reaching  inside  (site  II),  while  the  five  glucose  units  visible  for  the  second  molecules  (site  I),   between  three   domains:   the  N-­‐terminal,   the  carbohydrate   binding  module,   and  the   catalytic   domain.             161         ACTIVE  CENTER       Binding  site  I     Binding  site  II     Figure   V-­‐3:   Surface   depiction   of   RBEI   in   complex   with   M12.   At   the   center   of   this   groove,   residues  involved   in  catalysis  are  shown   in  blue.  These   catalytic  residues,   Y235,  D270,   H275,   R342,  D344,  E399,  H467  and  D468  according  to  RBEI  sequence  numbering,  (2)  were  predicted   based  on  biochemical  and  structural  data  of  ι-­‐amylase.  (3,  4)     V-­‐2-­‐1  Binding  Site  I:  Carbohydrate  Binding  Module  (CBM)       This   binding   site   has   been   observed   in   both   the   truncated   structures   of   RBEI,   with   maltopentaose   (M5)   and   maltododecaose   (M12).   Based   on   the   crystal   structure,   this   site   bridges   the   CBM48,   N   terminal   helical   and   catalytic   domains.   In   both   structures,   five   units   of   glucose  are  visible  in  this  site  I.  The  saccharides  make  several  hydrogen-­‐bond  interactions  with   the  residues  in  that  side  along  with  the  water  networks.  (Figure  V-­‐4).       The  current  rice  Branching  Enzyme  I  structure  confirms  once  again  the  necessity  of  the  CBM48   domain  for  carbohydrate  binding.  (TableV-­‐1).   162   Tr p 72 His 294 Glu 320 Glu Trp 320 72 Trp 319 Arg 323 Glu 45 Glu 295 3 His 44 Pro 74 4     Lys 97 2 Phe 100 5 Lys 99 1   Figure   V-­‐4:   Binding   site   I:   detailed   interactions   between   the   oligosaccharide   and   RBEI.   The   protein  atoms  are  colored  by  type:  C  in  blue  marine,  O  in  red  and  N  in  dark  blue.  M12  atoms  are   also   colored   by   type   with   C   in   yellow   and   O   in   red.   Glucose   units   are   numbered   in   red.   Hydrogen   bonds   are   shown   in   dotted   black   lines.   Water   molecules   interacting   in   this   site   are   represented  in  spheres  and  colored  in  cyan.   V-­‐2-­‐2  New  Observed  Binding  Site  II     The   structure   of   M12-­‐bound   RBEI   reveals   a   second   M12   molecule   bound.   The   second   molecule   of   oligosaccharide   and   close   residues   around   it,   are   divided   into   two   binding   sites.   Binding  site  II  are  starting  away  from  the  active  site.  (Figure  V-­‐2)   Almost  all  glucose  units  in  this  carbohydrate  make  interactions  (specially  hydrogen  binding)  to   residues  of  the  enzyme.  (Figure  V-­‐6  and  Table  V-­‐1)     163     Lys 484 Met 490 Thr 488 Met 486 Asp 483 5 Lys 475 6 10 8 12 Leu 562 His 561 Phe 479 7 9 11   Tyr 487 Ser 491 Leu 493 Pro 533 Tyr 564 Leu 556   3 4 Gln 553 Glu 534 Gly 473 2   1 Trp 535 Val 472 Ser 470 Asp 537 Phe 538 Arg 540 Figure  V-­‐5:  Binding  site  II,  detailed  interactions  between  the  oligosaccharide  and  RBEI.  Numbers   for   glucose   units,   and   the   protein   atoms   are   colored.   Hydrogen   bonds   and   water   molecules   are   also  represented.  Hydrogen  bonds  are  shown  in  dotted  black  lines.   V-­‐3  Mutational  studies  on  the  observed  binding  sites  and  active  site  of  RBEI   Several  essential  binding  sites  and  conserved  amino  acid  residues  are  now  revealed  in   RBEI.  Therefore  in  order  to  understand  the  effect  of  each  residue  on  the  protein  activity  and  its   involvement  in  determination  of  chain  length  transferred,  it  becomes  important  to  perform  a   series   of   mutations   starting   with   point   mutations,   and   then   after   comparing   results   carrying   double  mutants  of  RBEI.         164   Table  V-­‐1:  Residues  involved  in  the  interactions  in  Binding  sites  I,  II  and  III  of  RBEI-­‐M12.     Binding  site   Residues  involved  in  binding       I     H44,  E45,  W72,  P74,  K97,  K99,  F100,  W319,  H294,       E295,  E320,  R323       II   H561,  S491,  Y487,  T488,  M490,  Y564,  M486,  Q553,   E534,   L493,   L562,   L556,   P533,   K484,   D483,   K475,   G473,  S470,  D537,  F538,  R540,  F479,  W535,  V472             V-­‐3-­‐1  Effect  of  Active  Site  Mutation:       As   previously   explained   in   section   V-­‐2,   Asp344   is   the   nucleophile   in   the     catalytic   process.   (Figure   V-­‐3,  Figure  V-­‐6)   to   confirm   this   the   D344A   mutant  was   made   to   study   its   effect   on  the  enzyme  activity  and  function.  The  goal  is  to  confirm  its’  role  as  nucleophile,  and  to  use   the   in-­‐active   enzyme   in   crystallization   experiments,   because   the   inactive   enzyme   may   bind   a   glucan  in  the  active  site  without  degradation.   165         Active  site   Glu 399 Binding  site  I Asp 344           Binding  site  II   Figure   V-­‐6:   Overall   structure   of   RBEI.   Active   site   residues   are   shown   in   blue   sticks.   Two   oligosaccharide   molecules   are   in   yellow   sticks.   Oxygen   atoms   are   in   red.   So   far,   we   have   not   identified  any  glucose  units  close  and  around  the  active  site  yet.     The   iodine-­‐activity   assay   was   carried   out   on   purified   mutated   enzyme.   Please   refer   to   section   V-­‐4-­‐1   and   V-­‐4-­‐2   for   full   details   on   the   purification   step   and   the   assay   protocol.   The   absorbance  of  the  glucan–iodine  complex  was  almost  constant  during  this  time,  which  indicates   the  deactivation  of  enzyme  due  to  the  mutation  in  the  active  site.  (Figure  V-­‐7)  Specific  acitivy  of   D344A  was  0.02  Οmol/min-­‐1mg-­‐1  (U/mg)  which  was  only  0.7%  active  compared  to  the  activity  of   wild-­‐type   (WT)   truncated   RBEI.   Data   confirmed   that   we   knocked   out   the   enzyme   activity   by   mutating  its  active  site  as  we  predicited.     Crystallization  attempts  with  this  mutant  have  so  far  not  yielded  well-­‐diffracting  crystals.       166   Absorbance  at  660nm  (a.  u.)         D344A  Acmvity  Assay     1.6       1.4       1.2     1       0.8   D344A       y  =  -­‐0.0006x  +  1.4839     0.6       0.4       0.2     0     0   5   10   15   20   25   30   35     Time  (min)     Figure  V-­‐7:  Absorbance  versus  Time  for  the  D344A  mutation  of  truncated  RBEI.The  graph  is  the   average  for  three  trials.    Slope/Protein  indicates  the  activity  (U/mg)  of  the  enzyme.     V-­‐3-­‐1  Effect  of  E534A  Mutation:   In   binding   site   II   of   RBEI-­‐M12,   Glu   534   directly   interacts   with   Glucose7   via   a   hydrogen   bond  to  O2  and  makes  a  water-­‐mediated  interaction  with  O3  (Figure  V-­‐5,  Figure  V-­‐8).     167   10 7     9 11   12 6   8 Glu 534 Figure  V-­‐8:  Detailed  interactions  between  the  Glu7  in  binding  site  II  and  residue  E534  of  RBEI.   All   interactions   between   them   are   in   between   2-­‐3.5   Å.     Numbers   for   glucose   units,   and   the   protein   atoms   are   colored.   Hydrogen   bonds   and   water   molecules   are   also   represented.   Hydrogen  bonds  are  shown  in  dotted  black  lines.         The  E534A  mutation  was  made  and  assayed  to  test  the  importance  of  this  interaction,   which  is  located  far  from  the  active  site.    However  the  iodine-­‐activity  assay  result  surprised  us   and   E534A   mutant   was   83%   active   compare   to   the   WT   enzyme.   (Figure   V-­‐9)   Based   on   our   result,   Glu   534   in   Site   II,   is   more   likely   to   involve   binding   enzyme   to   its   substrate   than   directing   oligosaccharide   into   the   active   site   (because   mutation   of   this   residue,   cannot   deactivate   the   enzyme).    This  opens  up  the  idea  that  maybe  residues  around  that  site  are  most  likely  to  involve   directing   transferring   chains   (donor)   and   acceptor   chains   into   the   active   site.     To   confirm   our   hypothesis  regarding  binding  site  II,  a  list  of  series  of  mutations  were  proposed  (Table  V-­‐2).  The   168   mutation’s   DNA   were   prepared   but   they   need   to   be   expressed   and   tested   for   their   activity   and   crystallized  in  the  future.     E534A    Acmvity  Assay   Absorbance  at  660nm  (a.  u.  )   0.7   0.6   0.5   0.4   0.3   0.2   0.7   0.6   0.5   0.4   0.3   E534A   0.2   y  =  -­‐0.0729x  +  0.628     0.1   0     Absorbance  at  660nm  (a.  u.  )                                         0   2   4   Time  (min)   6   0.1   0   0   5   10   15   20   25   30   Time  (min)   Figure  V-­‐9:  Absorbance   versus   Time   for  the  E534A  mutation  of  truncated  RBEI:  The   graph  is   the   average   for   three   trials.   The   decrease   in   absorbance   is   linear   for   the   first   5   min   then   reaches  a  plateau  due  to  the  high  activity  of  the  enzyme.                                   169   Table  V-­‐2:     More  suggested  mutations  to  study  for  residues  interacting  with  M12  in  binding  site  II  of   RBEI.   ! ! Residue'Number!       ! ! Residue! ! Type%of% interaction) to#glucose( units! ! ! Mutation(! 561! His! Hydrophilic! Ala! 562! Leu! Hydrophobic! Phe! 564! Tyr! Hydrophilic! Phe! 487! Tyr! Hydrophilic! Ala!       V-­‐4  Experimental   V-­‐4-­‐1  Material  and  Method   Full-length RBEI (Oryza Sativa, Clone ID: 114619 from Japan) was cloned into a modified pet28a vector (Novagen) by Dr. Fawaz. It contained N-terminus His6-tag and SUMO site right before the N-terminus of RBEI (the first 65 residues were excluded). There is also a tobacco etch virus (TEV) recognition site for cleavage in the C-terminus of the RBEI sequence (between res 694-695). To mutate the WT DNA, following primers were ordered and site-directed mutagenesis was performed to mutate and result in pure   DNA.   This   method   was   performed   by   pfuTutrbo   DNA   polymerase   (agilent)   and   temperature  circle  in  PCR  instrument.  (Table  V-­‐3)     170   PCR  Protocol  for  RBEI  mutagenesis:     1x                        95°C,    3min       20x   95°C  ,    30  sec                 (Tm-­‐4)°C  ,    3  min                                                                   72  °C  ,    5min   1x                      72°C,  10  min     1x                    25°C,    5  min     Tm   stands   for   the   melting   temperature   of   the   primer   for   each   mutation   and   it   is   mentioned   on   each  ordered  primer  from  IDT.       PCR  Primers  for  D344A  mutant:   Forward:  5Ęź-­‐  GGCTTCCGATTTGCTGGGGTTACGTCA  -­‐3Ęź   Reverse:  5Ęź-­‐  TGATGTAACCCCAGCAAATCGGAAGCC  -­‐3Ęź   PCR  Primers  for  E534A  mutant:   Forward:  5Ęź-­‐  TTTGGCCATCCAGCCTGGATTGACTTT-­‐3Ęź   Reverse:  5Ęź-­‐  AAAGTCCAATCCAGGCTGGATGGCCAAA-­‐3Ęź   PCR  Primers  for  D344A  mutant:   Forward:  5Ęź-­‐  GGCTTCCGATTTGCTGGGGTTACGTCA  -­‐3Ęź   Reverse:  5Ęź-­‐  TGATGTAACCCCAGCAAATCGGAAGCC  -­‐3Ęź   PCR  Primers  for  E534A  mutant:   Forward:  5Ęź-­‐  TTTGGCCATCCAGCCTGGATTGACTTT-­‐3Ęź   Reverse:  5Ęź-­‐  AAAGTCCAATCCAGGCTGGATGGCCAAA-­‐3Ęź   PCR  Primers  for  H561A  mutant:   Forward:  5Ęź-­‐  CGACACTGATGCCCTTCGATACA-­‐3Ęź   Reverse:  5Ęź-­‐  TGTATCGAAGGGCATCAGTGTCG-­‐3Ęź   171   PCR  Primers  for  D344A  mutant:   Forward:  5Ęź-­‐  GGCTTCCGATTTGCTGGGGTTACGTCA  -­‐3Ęź   Reverse:  5Ęź-­‐  TGATGTAACCCCAGCAAATCGGAAGCC  -­‐3Ęź   PCR  Primers  for  E534A  mutant:   Forward:  5Ęź-­‐  TTTGGCCATCCAGCCTGGATTGACTTT-­‐3Ęź   Reverse:  5Ęź-­‐  AAAGTCCAATCCAGGCTGGATGGCCAAA-­‐3Ęź   PCR  Primers  for  D344A  mutant:   Forward:  5Ęź-­‐  GGCTTCCGATTTGCTGGGGTTACGTCA  -­‐3Ęź   Reverse:  5Ęź-­‐  TGATGTAACCCCAGCAAATCGGAAGCC  -­‐3Ęź   PCR  Primers  for  E534A  mutant:   Forward:  5Ęź-­‐  TTTGGCCATCCAGCCTGGATTGACTTT-­‐3Ęź   Reverse:  5Ęź-­‐  AAAGTCCAATCCAGGCTGGATGGCCAAA-­‐3Ęź   PCR  Primers  for  H561A  mutant:   Forward:  5Ęź-­‐  CGACACTGATGCCCTTCGATACA-­‐3Ęź   Reverse:  5Ęź-­‐  TGTATCGAAGGGCATCAGTGTCG-­‐3Ęź     PCR  Primers  for  H561A  mutant:   Forward:  5Ęź-­‐  CGACACTGATGCCCTTCGATACA-­‐3Ęź   Reverse:  5Ęź-­‐  TGTATCGAAGGGCATCAGTGTCG-­‐3Ęź     PCR  Primers  for  L562F  mutant:   Forward:  5Ęź-­‐  TACAATTCCTTGGCCATCAAGAGAA-­‐3Ęź   Reverse:  5Ęź-­‐  TTCTCTTGATGGCCAAGGAAATGTA-­‐3Ęź   PCR  Primers  for  Y564F  mutant:   Forward:  5Ęź-­‐  GCATTCATATACTTGAATCGAAGGTGATCAG-­‐3Ęź   Reverse:  5Ęź-­‐  CTGATCACCTTCGATTCAAGTATATGAATGC-­‐3Ęź   PCR  Primers  for  H561A  mutant:   Forward:  5Ęź-­‐  CGACACTGATGCCCTTCGATACA-­‐3Ęź   Reverse:  5Ęź-­‐  TGTATCGAAGGGCATCAGTGTCG-­‐3Ęź     PCR  Primers  for  L562F  mutant:   Forward:  5Ęź-­‐  TACAATTCCTTGGCCATCAAGAGAA-­‐3Ęź   Reverse:  5Ęź-­‐  TTCTCTTGATGGCCAAGGAAATGTA-­‐3Ęź   PCR  Primers  for  Y564F  mutant:   Forward:  5Ęź-­‐  GCATTCATATACTTGAATCGAAGGTGATCAG-­‐3Ęź   Reverse:  5Ęź-­‐  CTGATCACCTTCGATTCAAGTATATGAATGC-­‐3Ęź   172   Protein   over-­‐expression:   The   desired   DNA   of   RBEI   was   then   transformed   into   BL21(DE3)   E.coli   cells.   The   cells   were   grown   in   LB   media,   after   the   OD600   reached   too   0.5-­‐0.6   induced   with   0.5mM  IPTG  and  over  expressed  at  25°C  for  5  hours  before  centrifugation  and  freezing  at  -­‐20°C.       The   frozen   cells   (obtained   from   6L)   were   re-­‐suspended   and   supplemented   with   one   table   of   protease   inhibitor   tablet   in   lysis   buffer   (5mL/g   of   pellelts)   of   (50mM   Tris,   pH=8.0,   100mM   NaCl,   1mM   BME,   10mM   imidazole).   The   re-­‐suspended   cells   were   lysed   by   sonication   (three   1   min   with  1  min  relaxation  between  each)  and  centrifuged  for  20  min,  13000  rpm.  The  supernatant   was  subjected  to  Ni-­‐NTA  affinity  resin  (from  Qiagen).  Resin  washed  with  the  same  buffer  plus   30  mM  imidazole  until  it  reached  to  the  baseline.  Next,  Sumo  protease  added  on  the  column   before   eluting   the   protein   and   it   was   collected   in   different   fractions.   30   ÎźL   of   sample   is   collected  for  running  polyacrylamide  gel  electrophoresis  (SDS-­‐PAGE)  at  each  step.     The  C-­‐terminus  cut  of  RBEI  was  carried  by  addition  of  the  TEV  protease  (Concentration  2  mg/mL)   and  dialyzed  against  buffer  50mM  Tris-­‐HCL  pH  8.0,  0.5mM  EDTA,  1mM  DTT  for  overnight  at  4°C.   Protein  was  concentrated  using  centrifugal  concentrator  with  10000  MW  cutoffs  and  purified   on  a  Superdex  75  16/600  GE  column.  Fractions  with  right  size  of  truncated  RBEI  (80kDa)  were   analyzed   by   12%   SDS   PAGE.   Pure   protein   then   was   concentrated   by   centrifugal   concentrator   to   (3-­‐5  mg/ml),  as  determined  by  Bio-­‐Rad  protein  assay  at  600nm.   V-­‐4-­‐2  Activity  Assay   We   assayed   the   enzymatic   activity   for   the   truncated   RBEI   through   iodine-­‐staining   assay.   Branching   enzymes   have   been   tested   for   their   activity   since   the   1970s   (131,   154,   155).   The   most   commonly   used   assay,   was   developed   by   Boyer,   C.   and   Preiss,   J.   in   the   Biochemistry   department   of   Michigan   State   University   (MSU).   This   assay   is   based   on   the   decrease   in   173   absorption   of   a   glucan-­‐iodine   complex.   (154).   The   absorbance   of   the   glucan–iodine   complex   is   decreased   by   the   branching   of   the   substrate   (amylose   from   Sigma   Scientific)   with   branching   enzyme.  There  are  3  trials  for  this  assay  simultaneously  (to  minimize  variables  in  substrates  or   instrument   and   therefore   obtained   three   data   sets   per   protein)   all   contain   80𝜇L   of   amylose   stock  solution  (0.5ml  10%NaOH,  50mg  amylose,  2ml  H2O),  6  𝜇L  HCL  and  820  𝜇L  H2O.  Solutions   are  set  to  pH  8.0.  Tubes  stabilize  at  30°C  for  about  10  minutes.  The  iodine  reagent  was  made   daily   from   2.9   mL   of   stock   solution   (0.26   g   of   I2   and   2.6g   of   KI   in   10mL   of   water).   Branching   enzyme  I  (30  𝜇g)  is  added  to  each  tube  of  amylose  and  an  aliquot  of  50  ΟL  of  reaction  mixture  is   withdrawn   to   add   to   the   iodine   solution   every   5   minutes   over   a   30   minute-­‐period.   The   absorbance   was   determined   at   660   nm   for   the   reaction   mixtures   containing   amylose.   We   monitor  the  change  in  absorbance  (y-­‐axis)  at  a  wavelength  of  660nm  in  5  minute  intervals  (X-­‐ axis)  over  a  30-­‐minute  total  reaction  time.    The  greater  decrease  in  absorbance,  the  greater  the   activity.     Activity (U / mg) = Absorbance / min slope =   Protein Amount (mg) 0.03 mg Percent Activity = Protein Activity (U/mg) × 100   WT Activity (U/mg) In   the  case  of   E534A  Mutant,   the   plot  of   absorbance   versus   time   was   linear   for   the   first   5   min,   but   then   reached   a   plateau.   Therefore,   we   used   the   linear   part   of   the   graph   to   calculate   the  slope.  One  unit  of  activity  is  defined  as  a  decrease  in  absorbance  of  1.0  absorbance  unit  per   min  at  660  nm  (154)  and  is  measured  in  U/mg  of  protein.  (154,  156)     174   V-­‐4-­‐3  Crystallization  of  RBEI  mutants     Each  mutant  protein  was  concentrated  to  3-­‐5mg/mL.  Initial  screens  were  done  on  the   mutant’s   proteins.   After   optimizing   different   conditions   for   crystals   to   grow,   the   best   condition   was   28%   PEG   8000,   550mM   NaOAC,   0.1   M   (CH3)2AsO2Na   ·∙   3H2O   pH   6.9.   Few   small   crystals   formed  after  a  month.  However  they  still  were  small  and  did  not  diffract.                                             175                 REFERENCES                                         176   REFERENCES         1.     2.     3.     4.     5.     6.     7.     8.     9.     10.   Harder,  M.  E.,  Deinzer,  M.  L.,  Leid,  M.  E.,  and  Schimerlik,  M.  I.  (2004)  Global  analysis  of   three-­‐state  protein  unfolding  data,  Protein  Science  13,  2207-­‐2222.   Noguchi,  J.,  Chaen,  K.,  Vu,  N.  T.,  Akasaka,  T.,  Shimada,  H.,  Nakashima,  T.,  Nishi,  A.,  Satoh,   H.,   Omori,   T.,   Kakuta,   Y.,   and   Kimura,   M.   (2011)   Crystal   structure   of   the   branching   enzyme  I  (BEI)  from  Oryza  sativa  L  with  implication  sfor  catalysis  and  susbtrate  binding,   Glycobiology  21,  1108-­‐1116.   Brzozowski,   A.   M.,   and   Davies,   G.   J.   (1997)   Structure   of   the   Aspergillus   oryzae   Alpha-­‐ Amylase   Complexed   with   the   Inhibitor   Acarbose   at   2.0   Å   Resolution,   Biochemistry   36,   10837-­‐10845.   Uitdehaag,  J.  C.  M.,  van  Alebeek,  G.  J.,  Van  Der  Veen,  B.,  Dijkhuizen,  L.,  and  Dijkstra,  B.   W.  (2000)  Structures  of  Maltohexaose  and  Maltoheptaose  Bound  at  the  Donor  Sites  of   Cyclodextrin  Glycosyltransferase  Give  Insight  into  the  Mechanisms  of  Transglycosylation   Activity  and  Cyclodextrin  Size  Specificity,  Biochemistry  39,  7772-­‐7780.   Ong,  D.  E.  (1984)  A  NOVEL  RETINOL-­‐BINDING  PROTEIN  FROM  RAT  -­‐  PURIFICATION  AND   PARTIAL  CHARACTERIZATION,  Journal  of  Biological  Chemistry  259,  1476-­‐1482.   Nossoni,   Z.,   Assar,   Z.,   Yapici,   I.,   Nosrati,   M.,   Wang,   W.,   Berbasova,   T.,   Vasileiou,   C.,   Borhan,   B.,   and   Geiger,   J.   (2014)   Structures   of   holo   wild-­‐type   human   cellular   retinol-­‐ binding  protein  II  (hCRBPII)  bound  to  retinol  and  retinal,  Acta  crystallographica.  Section   D,  Biological  crystallography  70,  3226-­‐3232.   Veerkamp,   J.   H.,   Vankuppevelt,   T.,   Maatman,   R.,   and   Prinsen,   C.   F.   M.   (1993)   STRUCTURAL   AND   FUNCTIONAL-­‐ASPECTS   OF   CYTOSOLIC   FATTY-­‐ACID-­‐BINDING   PROTEINS,  Prostaglandins  Leukotrienes  and  Essential  Fatty  Acids  49,  887-­‐906.   Banaszak,   L.,   Winter,   N.,   Xu,   Z.   H.,   Bernlohr,   D.   A.,   Cowan,   S.,   and   Jones,   T.   A.   (1994)   LIPID-­‐BINDING   PROTEINS   -­‐   A   FAMILY   OF   FATTY-­‐ACID   AND   RETINOID   TRANSPORT   PROTEINS,  Advances  in  Protein  Chemistry,  Vol  45  45,  89-­‐151.   Vaezeslami,   S.,   Mathes,   E.,   Vasilelou,   C.,   Borhan,   B.,   and   Geiger,   J.   H.   (2006)   The   structure   of   apo-­‐wild-­‐type   cellular   retinoic   acid   binding   protein   II   at   1.4   A   and   its   relationship   to   ligand   binding   and   nuclear   translocation,   Journal   of   Molecular   Biology   363,  687-­‐701.   Haunerland,   N.   H.,   and   Spener,   F.   (2004)   Fatty   acid-­‐binding   proteins-­‐-­‐insights   from   genetic  manipulations,  Progress  in  lipid  research  43,  328-­‐349.   177   11.     12.     13.     14.     15.     16.     17.     18.     19.     20.   Lucke,   C.,   Huang,   S.,   Rademacher,   M.,   and   Ruterjans,   H.   (2002)   New   insights   into   intracellular   lipid   binding   proteins:   The   role   of   buried   water,   Protein   science   :   a   publication  of  the  Protein  Society  11,  2382-­‐2392.   Schaap,   F.   G.,   van   der   Vusse,   G.   J.,   and   Glatz,   J.   F.   (2002)   Evolution   of   the   family   of   intracellular   lipid   binding   proteins   in   vertebrates,   Molecular   and   cellular   biochemistry   239,  69-­‐77.   Smathers,   R.   L.,   and   Petersen,   D.   R.   (2011)   The   human   fatty   acid-­‐binding   protein   family:   evolutionary  divergences  and  functions,  Human  genomics  5,  170-­‐191.   Franzoni,   L.,   Lucke,   C.,   Perez,   C.,   Cavazzini,   D.,   Rademacher,   M.,   Ludwig,   C.,   Spisni,   A.,   Rossi,  G.  L.,  and  Ruterjans,  H.  (2002)  Structure  and  backbone  dynamics  of  apo-­‐  and  holo-­‐ cellular   retinol-­‐binding   protein   in   solution,   Journal   of   Biological   Chemistry   277,   21983-­‐ 21997.   Crow,   J.   A.,   and   Ong,   D.   E.   (1985)   CELL-­‐SPECIFIC   IMMUNOHISTOCHEMICAL   LOCALIZATION   OF   A   CELLULAR   RETINOL-­‐BINDING   PROTEIN   (TYPE-­‐2)   IN   THE   SMALL-­‐ INTESTINE  OF  RAT,  Proceedings  of  the  National  Academy  of  Sciences  of  the  United  States   of  America  82,  4707-­‐4711.   Vogel,  S.,  Mendelsohn,  C.  L.,  Mertz,  J.  R.,  Piantedosi,  R.,  Waldburger,  C.,  Gottesman,  M.   E.,  and  Blaner,  W.  S.  (2001)  Characterization  of  a  new  member  of  the  fatty  acid-­‐binding   protein   family   that   binds   all-­‐trans-­‐retinol,   Journal   of   Biological   Chemistry   276,   1353-­‐ 1360.   Folli,   C.,   Calderone,   V.,   Ottonello,   S.,   Bolchi,   A.,   Zanotti,   G.,   Stoppini,   M.,   and   Berni,   R.   (2001)   Identification,   retinoid   binding,   and   x-­‐ray   analysis   of   a   human   retinol-­‐binding   protein,   Proceedings   of   the   National   Academy   of   Sciences   of   the   United   States   of   America  98,  3710-­‐3715.   Li,  E.,  Demmer,  L.  A.,  Sweetser,  D.  A.,  Ong,  D.  E.,  and  Gordon,  J.  I.  (1986)  ISOLATION  OF  A   CLONED   CDNA-­‐ENCODING   AN   ABUNDANT   INTESTINAL   RETINOL   BINDING-­‐PROTEIN,   Clinical  Research  34,  A685-­‐A685.   Li,   E.,   and   Norris,   A.   W.   (1996)   Structure/function   of   cytoplasmic   vitamin   A-­‐binding   proteins,  Annual  Review  of  Nutrition  16,  205-­‐234.   Li,   E.,   Qian,   S.   J.,   Winter,   N.   S.,   Davignon,   A.,   Levin,   M.   S.,   and   Gordon,   J.   I.   (1991)   FLUORINE   NUCLEAR-­‐MAGNETIC-­‐RESONANCE   ANALYSIS   OF   THE   LIGAND-­‐BINDING   PROPERTIES   OF   2   HOMOLOGOUS   RAT   CELLULAR   RETINOL-­‐BINDING   PROTEINS   EXPRESSED  IN  ESCHERICHIA-­‐COLI,  Journal  of  Biological  Chemistry  266,  3622-­‐3629.   178   21.     22.     23.     24.     25.     26.     27.     28.   29.     30.     31.   Calderone,   V.,   Folli,   C.,   Marchesani,   A.,   Berni,   R.,   and   Zanotti,   G.   (2002)   Identification   and   structural   analysis   of   a   zebrafish   apo   and   holo   cellular   retinol-­‐binding   protein,   Journal  of  Molecular  Biology  321,  527-­‐535.   Tarter,  M.,  Capaldi,  S.,  Carrizo,  M.  E.,  Ambrosi,  E.,  Perduca,  M.,  and  Monaco,  H.  L.  (2008)   Crystal   structure   of   human   cellular   retinol-­‐binding   protein   11   to   1.2   angstrom   resolution,  Proteins-­‐Structure  Function  and  Bioinformatics  70,  1626-­‐1630.   Winter,   N.   S.,   Bratt,   J.   M.,   and   Banaszak,   L.   J.   (1993)   CRYSTAL-­‐STRUCTURES   OF   HOLO   AND   APO-­‐CELLULAR   RETINOL-­‐BINDING   PROTEIN-­‐II,   Journal   of   Molecular   Biology   230,   1247-­‐1259.   Kane,  M.  A.,  Bright,  F.  V.,  and  Napoli,  J.  L.  (2011)  Binding  affinities  of  CRBPI  and  CRBPII   for  9-­‐cis-­‐retinoids,  Biochimica  Et  Biophysica  Acta-­‐General  Subjects  1810,  514-­‐518.   Malpeli,   G.,   Stoppini,   M.,   Zapponi,   M.   C.,   Folli,   C.,   and   Berni,   R.   (1995)   INTERACTIONS   WITH   RETINOL   AND   RETINOIDS   OF   BOVINE   CELLULAR   RETINOL-­‐BINDING   PROTEIN,   European  Journal  of  Biochemistry  229,  486-­‐493.   McBee,   J.   K.,   Kuksa,   V.,   Alvarez,   R.,   de   Lera,   A.   R.,   Prezhdo,   O.,   Haeseleer,   F.,   Sokal,   I.,   and   Palczewski,   K.   (2000)   Isomerization   of   all-­‐trans-­‐retinol   to   cis-­‐retinols   in   bovine   retinal   pigment   epithelial   cells:   Dependence   on   the   specificity   of   retinoid-­‐binding   proteins,  Biochemistry  39,  11370-­‐11380.   Borshchevskiy,  V.  I.,  Round,  E.  S.,  Popov,  A.  N.,  Bueldt,  G.,  and  Gordeliy,  V.  I.  (2011)  X-­‐ ray-­‐Radiation-­‐Induced   Changes   in   Bacteriorhodopsin   Structure,   Journal   of   Molecular   Biology  409,  813-­‐825.   Matsui,  Y.,  Sakai,  K.,  Murakami,  M.,  Shiro,  Y.,  Adachi,  S.,  Okumura,  H.,  and  Kouyama,  T.   (2002)  Specific  damage  induced  by  X-­‐ray  radiation  and  structural  changes  in  the  primary   photoreaction  of  bacteriorhodopsin,  Journal  of  Molecular  Biology  324,  469-­‐481.   Schreckenbach,   T.,   Walckhoff,   B.,   and   Oesterhelt,   D.   (1977)   STUDIES   ON   RETINAL-­‐ PROTEIN  INTERACTION  IN  BACTERIORHODOPSIN,  European  Journal  of  Biochemistry  76,   499-­‐511.   Cowan,   S.   W.,   Newcomer,   M.   E.,   and   Jones,   T.   A.   (1993)   CRYSTALLOGRAPHIC   STUDIES   ON   A   FAMILY   OF   CELLULAR   LIPOPHILIC   TRANSPORT   PROTEINS   -­‐   REFINEMENT   OF   P2   MYELIN   PROTEIN   AND   THE   STRUCTURE   DETERMINATION   AND   REFINEMENT   OF   CELLULAR   RETINOL-­‐BINDING   PROTEIN   IN   COMPLEX   WITH   ALL-­‐TRANS-­‐RETINOL,   Journal   of  Molecular  Biology  230,  1225-­‐1246.   Vasileiou,   C.,   Vaezeslami,   S.,   Crist,   R.   M.,   Rabago-­‐Smith,   M.,   Geiger,   J.   H.,   and   Borhan,   B.   (2007)   Protein   design:   reengineering   cellular   retinoic   acid   binding   protein   II   into   a   rhodopsin  protein  mimic,  J  Am  Chem  Soc  129,  6140-­‐6148.   179   32.     33.     34.     35.     36.     37.     38.     39.     40.     41.     42.     43.     44.     Otwinowski,   Z.,   and   Minor,   W.   (1997)   Processing   of   X-­‐ray   diffraction   data   collected   in   oscillation  mode,  Macromolecular  Crystallography,  Pt  A  276,  307-­‐326.   Vagin,   A.,   and   Teplyakov,   A.   (1997)   MOLREP:   an   automated   program   for   molecular   replacement,  Journal  of  Applied  Crystallography  30,  1022-­‐1025.   Vagin,   A.,   and   Teplyakov,   A.   (2010)   Molecular   replacement   with   MOLREP,   Acta   Crystallographica  Section  D-­‐Biological  Crystallography  66,  22-­‐25.   Murshudov,   G.   N.,   Skubak,   P.,   Lebedev,   A.   A.,   Pannu,   N.   S.,   Steiner,   R.   A.,   Nicholls,   R.   A.,   Winn,   M.   D.,   Long,   F.,   and   Vagin,   A.   A.   (2011)   REFMAC5   for   the   refinement   of   macromolecular   crystal   structures,   Acta   Crystallographica   Section   D-­‐Biological   Crystallography  67,  355-­‐367.   Emsley,   P.,   Lohkamp,   B.,   Scott,   W.   G.,   and   Cowtan,   K.   (2010)   Features   and   development   of  Coot,  Acta  Crystallographica  Section  D-­‐Biological  Crystallography  66,  486-­‐501.   Bennett,   M.   J.,   Schlunegger,   M.   P.,   and   Eisenberg,   D.   (1995)   3D   domain   swapping   -­‐   A   mechanism  for  oligomer  assembly,  Protein  Science  4,  2455-­‐2468.   Crestfield,   A.   M.,   Stein,   W.   H.,   and   Moore,   S.   (1962)   On   the   aggregation   of   bovine   pancreatic  ribonuclease,  Archives  of  biochemistry  and  biophysics  Suppl  1,  217-­‐222.   Liu,   Y.,   and   Eisenberg,   D.   (2002)   3D   domain   swapping:   As   domains   continue   to   swap,   Protein  Science  11,  1285-­‐1299.   Liu,   Y.   S.,   Gotte,   G.,   Libonati,   M.,   and   Eisenberg,   D.   (2002)   Structures   of   the   two   3D   domain-­‐swapped  RNase  A  trimers,  Protein  Science  11,  371-­‐380.   Liu,   Y.   S.,   Gotte,   G.,   Libonati,   M.,   and   Eisenberg,   D.   S.   (2001)   Crystal   structure   of   an   Rnase  A  dimer  displaying  a  new  type  of  3D  domain  swapping,  Nucleic  Acid  Database.   Liu,  Y.  S.,  Hart,  P.  J.,  Schlunegger,  M.  P.,  and  Eisenberg,  D.  (1998)  The  crystal  structure  of   a   3D   domain-­‐swapped   dimer   of   RNase   A   at   a   2.1-­‐angstrom   resolution,   Proceedings   of   the  National  Academy  of  Sciences  of  the  United  States  of  America  95,  3437-­‐3442.   Liu,   Y.,   and   Eisenberg,   D.   (2002)   3D   domain   swapping:   as   domains   continue   to   swap,   Protein  science  :  a  publication  of  the  Protein  Society  11,  1285-­‐1299.   Koharudin,   L.   M.,   Liu,   L.,   and   Gronenborn,   A.   M.   (2013)   Different   3D   domain-­‐swapped   oligomeric   cyanovirin-­‐N   structures   suggest   trapped   folding   intermediates,   Proc   Natl   Acad  Sci  U  S  A  110,  7702-­‐7707.   180   45.     46.     47.     48.     49.     50.     51.     52.     53.     54.     55.     Assar,  Z.,  Nossoni,  Z.,  Wang,  W.,  Santos,  E.  M.,  Kramer,  K.,  McCornack,  C.,  Vasileiou,  C.,   Borhan,   B.,   and   Geiger,   J.   H.   (2016)   Domain-­‐Swapped   Dimers   of   Intracellular   Lipid-­‐ Binding  Proteins:  Evidence  for  Ordered  Folding  Intermediates,  Structure  24,  1590-­‐1598.   Bax,  B.,  Lapatto,  R.,  Nalini,  V.,  Driessen,  H.,  Lindley,  P.  F.,  Mahadevan,  D.,  Blundell,  T.  L.,   and   Slingsby,   C.   (1990)   X-­‐ray   analysis   of   beta-­‐B2-­‐crystallin   and   evolution   of   oligomeric   lens  proteins,  Nature  347,  776-­‐780.   Hakansson,  M.,  Svensson,  A.,  Fast,  J.,  and  Linse,  S.  (2001)  An  extended  hydrophobic  core   induces  EF-­‐hand  swapping,  Protein  Science  10,  927-­‐933.   Rousseau,   F.,   Schymkowitz,   J.,   and   Itzhaki,   L.   S.   (2012)   Implications   of   3D   domain   swapping   for   protein   folding,   misfolding   and   function,   In   Protein   Dimerization   and   Oligomerization  in  Biology  (Mattews,  J.  M.,  Ed.),  pp  137-­‐152.   Yang,   F.,   Bewley,   C.   A.,   Louis,   J.   M.,   Gustafson,   K.   R.,   Boyd,   M.   R.,   Gronenborn,   A.   M.,   Clore,  G.  M.,  and  Wlodawer,  A.  (1999)  Crystal  structure  of  cyanovirin-­‐N,  a  potent  HIV-­‐ inactivating  protein,  shows  unexpected  domain  swapping,  Journal  of  Molecular  Biology   288,  403-­‐412.   Szilagyi,   A.,   Zhang,   Y.,   and   Zavodszky,   P.   (2012)   Intra-­‐Chain   3D   Segment   Swapping   Spawns  the  Evolution  of  New  Multidomain  Protein  Architectures,  Journal  of  Molecular   Biology  415,  221-­‐235.   Stroud,   J.   C.,   Liu,   C.,   Teng,   P.   K.,   and   Eisenberg,   D.   (2012)   Toxic   fibrillar   oligomers   of   amyloid-­‐beta   have   cross-­‐beta   structure,   Proceedings   of   the   National   Academy   of   Sciences  of  the  United  States  of  America  109,  7717-­‐7722.   Cleary,   J.   P.,   Walsh,   D.   M.,   Hofmeister,   J.   J.,   Shankar,   G.   M.,   Kuskowski,   M.   A.,   Selkoe,   D.   J.,   and   Ashe,   K.   H.   (2005)   Natural   oligomers   of   the   amyloid-­‐protein   specifically   disrupt   cognitive  function,  Nature  Neuroscience  8,  79-­‐84.   Fandrich,   M.,   Fletcher,   M.   A.,   and   Dobson,   C.   M.   (2001)   Amyloid   fibrils   from   muscle   myoglobin   -­‐   Even   an   ordinary   globular   protein   can   assume   a   rogue   guise   if   conditions   are  right,  Nature  410,  165-­‐166.   Gronenborn,   A.   M.   (2009)   Protein   acrobatics   in   pairs   -­‐   dimerization   via   domain   swapping,  Current  Opinion  in  Structural  Biology  19,  39-­‐49.   Schlunegger,  M.  P.,  Bennett,  M.  J.,  and  Eisenberg,  D.  (1997)  Oligomer  formation  by  3D   domain   swapping:   A   model   for   protein   assembly   and   misassembly,   In   Advances   in   Protein  Chemistry,  Vol  50:  Protein  Misassembly  (Wetzel,  R.,  Ed.),  pp  61-­‐122.   181   56.     57.     58.     59.     60.     61.     62.     63.     64.     65.     66.     67.     68.   Selkoe,  D.  J.  (2004)  Folding  proteins  in  fatal  ways  (vol  426,  pg  900,  2003),  Nature  428,   445-­‐445.   Shtilerman,  M.  D.,  Ding,  T.  T.,  and  Lansbury,  P.  T.  (2002)  Molecular  crowding  accelerates   fibrillization   of   alpha-­‐synuclein:   Could   an   increase   in   the   cytoplasmic   protein   concentration  induce  Parkinson's  disease?,  Biochemistry  41,  3855-­‐3860.   Sunde,  M.,  and  Blake,  C.  (1997)  The  structure  of  amyloid  fibrils  by  electron  microscopy   and   X-­‐ray   diffraction,   In   Advances   in   Protein   Chemistry,   Vol   50:   Protein   Misassembly   (Wetzel,  R.,  Ed.),  pp  123-­‐159.   Bennett,  M.  J.,  Choe,  S.,  and  Eisenberg,  D.  (1994)  Refined  structure  of  dimeric  diphtheria   toxin  at  2.0  A  resolution,  Protein  science  :  a  publication  of  the  Protein  Society  3,  1444-­‐ 1463.   O'Neill,   J.   W.,   Kim,   D.   E.,   Johnsen,   K.,   Baker,   D.,   and   Zhang,   K.   Y.   (2001)   Single-­‐site   mutations   induce   3D   domain   swapping   in   the   B1   domain   of   protein   L   from   Peptostreptococcus  magnus,  Structure  9,  1017-­‐1027.   Ellis,   R.   J.   (2001)   Macromolecular   crowding:   an   important   but   neglected   aspect   of   the   intracellular  environment,  Curr  Opin  Struct  Biol  11,  114-­‐119.   Bennett,   M.   J.,   Choe,   S.,   and   Eisenberg,   D.   (1994)   Rlefined   strucutre   of   dimeric   diphtheria-­‐toxin  at  2.0-­‐angstrom  resolution,  Protein  Science  3,  1444-­‐1463.   Carroll,   S.   F.,   Barbieri,   J.   T.,   and   Collier,   R.   J.   (1986)   Dimeric   form   of   diphtheria-­‐toxin-­‐ purification  and  characterization,  Biochemistry  25,  2425-­‐2430.   Ellis,   R.   J.   (2001)   Macromolecular   crowding:   an   important   but   neglected   aspect   of   the   intracellular   environment   (vol   11,   pg   114,   2001),   Current   Opinion   in   Structural   Biology   11,  500-­‐500.   Minton,   A.   P.   (2000)   Implications   of   macromolecular   crowding   for   protein   assembly,   Current  Opinion  in  Structural  Biology  10,  34-­‐39.   O'Neill,   J.   W.,   Kim,   D.   E.,   Johnsen,   K.,   Baker,   D.,   and   Zhang,   K.   Y.   J.   (2001)   Single-­‐site   mutations   induce   3D   domain   swapping   in   the   B1   domain   of   protein   L   from   Peptostreptococcus  magnus,  Structure  9,  1017-­‐1027.   Gershenson,  A.,  Gierasch,  L.  M.,  Pastore,  A.,  and  Radford,  S.  E.  (2014)  Energy  landscapes   of  functional  proteins  are  inherently  risky,  Nature  chemical  biology  10,  884-­‐891.   Ignatova,  Z.,  and  Gierasch,  L.  M.  (2004)  Monitoring  protein  stability  and  aggregation  in   vivo  by  real-­‐time  fluorescent  labeling,  Proc  Natl  Acad  Sci  U  S  A  101,  523-­‐528.   182   69.     70.     71.     72.     73.     74.     75.     76.     77.     78.     79.     80.   Ignatova,  Z.,  and  Gierasch,  L.  M.  (2006)  Inhibition  of  protein  aggregation  in  vitro  and  in   vivo  by  a  natural  osmoprotectant,  Proc  Natl  Acad  Sci  U  S  A  103,  13357-­‐13361.   Ropson,  I.  J.,  Boyer,  J.  A.,  Schaeffer,  B.  A.,  and  Dalessio,  P.  M.  (2009)  Comparison  of  the   folding   mechanism   of   highly   homologous   proteins   in   the   lipid-­‐binding   protein   family,   Proteins:  Structure,  Function,  and  Bioinformatics  75,  799-­‐806.   Burns,   L.   L.,   and   Ropson,   I.   J.   (2001)   Folding   of   intracellular   retinol   and   retinoic   acid   binding  proteins,  Proteins:  Structure,  Function,  and  Bioinformatics  43,  292-­‐302.   Gershenson,  A.,  Gierasch,  L.  M.,  Pastore,  A.,  and  Radford,  S.  E.  (2014)  Energy  landscapes   of  functional  proteins  are  inherently  risky,  Nat.  Chem.  Biol.  10,  884-­‐891.   Crist,   R.   M.,   Vasileiou,   C.,   Rabago-­‐Smith,   M.,   Geiger,   J.   H.,   and   Borhan,   B.   (2006)   Engineering  a  Rhodopsin  Protein  Mimic,  Journal  of  the  American  Chemical  Society  128,   4522-­‐4523.   Berbasova,  T.,  Nosrati,  M.,  Vasileiou,  C.,  Wang,  W.,  Lee,  K.  S.,  Yapici,  I.,  Geiger,  J.  H.,  and   Borhan,  B.  (2013)  Rational  design  of  a  colorimetric  pH  sensor  from  a  soluble  retinoic  acid   chaperone,  J  Am  Chem  Soc  135,  16111-­‐16119.   Wang,  W.,  Geiger,  J.  H.,  and  Borhan,  B.  (2014)  The  photochemical  determinants  of  color   vision:   revealing   how   opsins   tune   their   chromophore's   absorption   wavelength,   BioEssays  :  news  and  reviews  in  molecular,  cellular  and  developmental  biology  36,  65-­‐74.   Wang,   W.,   Nossoni,   Z.,   Berbasova,   T.,   Watson,   C.   T.,   Yapici,   I.,   Lee,   K.   S.,   Vasileiou,   C.,   Geiger,   J.   H.,   and   Borhan,   B.   (2012)   Tuning   the   electronic   absorption   of   protein-­‐ embedded  all-­‐trans-­‐retinal,  Science  338,  1340-­‐1343.   Nosrati,   M.,   Berbasova,   T.,   Vasileiou,   C.,   Borhan,   B.,   and   Geiger,   J.   H.   (2016)   A   Photoisomerizing  Rhodopsin  Mimic  Observed  at  Atomic  Resolution,  J  Am  Chem  Soc  138,   8802-­‐8808.   Berbasova,   T.,   Santos,   E.   M.,   Nosrati,   M.,   Vasileiou,   C.,   Geiger,   J.   H.,   and   Borhan,   B.   (2016)  Light-­‐Activated  Reversible  Imine  Isomerization:  Towards  a  Photochromic  Protein   Switch,  Chembiochem  :  a  European  journal  of  chemical  biology  17,  407-­‐414.   Yapici,   I.,   Lee,   K.   S.,   Berbasova,   T.,   Nosrati,   M.,   Jia,   X.,   Vasileiou,   C.,   Wang,   W.,   Santos,   E.   M.,  Geiger,  J.  H.,  and  Borhan,  B.  (2015)  "Turn-­‐on"  protein  fluorescence:  in  situ  formation   of  cyanine  dyes,  J  Am  Chem  Soc  137,  1073-­‐1080.   Wang,  W.,  Nossoni,  Z.,  Berbasova,  T.,  Watson,  C.  T.,  Yapici,  I.,  Lee,  K.  S.  S.,  Vasileiou,  C.,   Geiger,   J.   H.,   and   Borhan,   B.   (2012)   Tuning   the   Electronic   Absorption   of   Protein-­‐ Embedded  All-­‐trans-­‐Retinal,  Science  338,  1340-­‐1343.   183   81.     82.     83.     84.     85.     86.     87.     88.     89.     90.     91.   Gunasekaran,   K.,   Hagler,   A.   T.,   and   Gierasch,   L.   M.   (2004)   Sequence   and   structural   analysis   of   cellular   retinoic   acid-­‐binding   proteins   reveals   a   network   of   conserved   hydrophobic  interactions,  Proteins-­‐Structure  Function  and  Genetics  54,  179-­‐194.   Nossoni,   Z.,   Assar,   Z.,   Yapici,   I.,   Nosrati,   M.,   Wang,   W.,   Berbasova,   T.,   Vasileiou,   C.,   Borhan,   B.,   and   Geiger,   J.   (2014)   Structures   of   holo   wild-­‐type   human   cellular   retinol-­‐ binding  protein  II  (hCRBPII)  bound  to  retinol  and  retinal,  Acta  Crystallogr.  Sect.  D.  Biol.   Crystallogr.  70,  3226-­‐3232.   Shameer,   K.,   Shingate,   P.   N.,   Manjunath,   S.   C.   P.,   Karthika,   M.,   Pugalenthi,   G.,   and   Sowdhamini,   R.   (2011)   3DSwap:   curated   knowledgebase   of   proteins   involved   in   3D   domain  swapping,  Database-­‐the  Journal  of  Biological  Databases  and  Curation.   Sillitoe,   I.,   Cuff,   A.   L.,   Dessailly,   B.   H.,   Dawson,   N.   L.,   Furnham,   N.,   Lee,   D.,   Lees,   J.   G.,   Lewis,   T.   E.,   Studer,   R.   A.,   Rentzsch,   R.,   Yeats,   C.,   Thornton,   J.   M.,   and   Orengo,   C.   A.   (2013)  New  functional  families  (FunFams)  in  CATH  to  improve  the  mapping  of  conserved   functional  sites  to  3D  structures,  Nucleic  Acids  Research  41,  D490-­‐D498.   Englander,   S.   W.,   and   Mayne,   L.   (2014)   The   nature   of   protein   folding   pathways,   Proc   Natl  Acad  Sci  U  S  A  111,  15873-­‐15880.   Englander,  S.  W.,  Mayne,  L.,  and  Krishna,  M.  M.  (2007)  Protein  folding  and  misfolding:   mechanism  and  principles,  Quarterly  reviews  of  biophysics  40,  287-­‐326.   Krishna,   M.   M.,   and   Englander,   S.   W.   (2007)   A   unified   mechanism   for   protein   folding:   predetermined   pathways   with   optional   errors,   Protein   science   :   a   publication   of   the   Protein  Society  16,  449-­‐464.   Mayne,   L.,   and   Englander,   S.   W.   (2000)   Two-­‐state   vs.   multistate   protein   unfolding   studied   by   optical   melting   and   hydrogen   exchange,   Protein   science   :   a   publication   of   the   Protein  Society  9,  1873-­‐1877.   Roder,  H.,  Elove,  G.  A.,  and  Englander,  S.  W.  (1988)  Structural  characterization  of  folding   intermediates   in   cytochrome   c   by   H-­‐exchange   labelling   and   proton   NMR,   Nature   335,   700-­‐704.   Martins,   N.   F.,   and   Santoro,   M.   M.   (1999)   Partially   folded   intermediates   during   trypsinogen  denaturation,  Brazilian  journal  of  medical  and  biological  research  =  Revista   brasileira  de  pesquisas  medicas  e  biologicas  32,  673-­‐682.   Medina,   E.,   Cordova,   C.,   Villalobos,   P.,   Reyes,   J.,   Komives,   E.   A.,   Ramirez-­‐Sarmiento,   C.   A.,   and   Babul,   J.   (2016)   Three-­‐Dimensional   Domain   Swapping   Changes   the   Folding   Mechanism  of  the  Forkhead  Domain  of  FoxP1,  Biophysical  journal  110,  2349-­‐2360.     184   92.     93.     94.     95.     96.     97.     98.     99.     100.     101.     102.   Zwanzig,   R.   (1997)   Two-­‐state   models   of   protein   folding   kinetics,   Proceedings   of   the   National  Academy  of  Sciences  of  the  United  States  of  America  94,  148-­‐150.   Sanson,   B.,   Wang,   T.,   Sun,   J.,   Wang,   L.,   Kaczocha,   M.,   Ojima,   I.,   Deutsch,   D.,   and   Li,   H.   (2014)  Crystallographic  study  of  FABP5  as  an  intracellular  endocannabinoid  transporter,   Acta  crystallographica.  Section  D,  Biological  crystallography  70,  290-­‐298.   Assar,   Z.,   Nossoni,   Z.,   Wang,   W.,   Santos,   Elizabeth  M.,   Kramer,   K.,   McCornack,   C.,   Vasileiou,  C.,  Borhan,  B.,  and  Geiger,  James  H.  Domain-­‐Swapped  Dimers  of  Intracellular   Lipid-­‐Binding  Proteins:  Evidence  for  Ordered  Folding  Intermediates,  Structure.   Silvaroli,   J.   A.,   Arne,   J.   M.,   Chelstowska,   S.,   Kiser,   P.   D.,   Banerjee,   S.,   and   Golczak,   M.   (2016)   Ligand   Binding   Induces   Conformational   Changes   in   Human   Cellular   Retinol-­‐ binding  Protein  1  (CRBP1)  Revealed  by  Atomic  Resolution  Crystal  Structures,  The  Journal   of  biological  chemistry  291,  8528-­‐8540.   Gillilan,  R.  E.,  Ayers,  S.  D.,  and  Noy,  N.  (2007)  Structural  basis  for  activation  of  fatty  acid-­‐ binding  protein  4,  J  Mol  Biol  372,  1246-­‐1260.   Sessler,  R.  J.,  and  Noy,  N.  (2005)  A  ligand-­‐activated  nuclear  localization  signal  in  cellular   retinoic  acid  binding  protein-­‐II,  Molecular  cell  18,  343-­‐353.   Armstrong,  E.  H.,  Goswami,  D.,  Griffin,  P.  R.,  Noy,  N.,  and  Ortlund,  E.  A.  (2014)  Structural   Basis   for   Ligand   Regulation   of   the   Fatty   Acid-­‐binding   Protein   5,   Peroxisome   Proliferator-­‐ activated   Receptor   β/δ   (FABP5-­‐PPARβ/δ)   Signaling   Pathway,   The   Journal   of   biological   chemistry  289,  14941-­‐14954.   Greenfield,  N.  J.  (2006)  Using  circular  dichroism  collected  as  a  function  of  temperature   to  determine  the  thermodynamics  of  protein  unfolding  and  binding  interactions,  Nature   Protocols  1,  2527-­‐2535.   Ramos,   C.   R.   R.,   Oyama,   S.,   Jr.,   Sforca,   M.   L.,   Pertinhez,   T.   A.,   Ho,   P.   L.,   and   Spisni,   A.   (2008)   2POA:   Schistosoma   mansoni   Sm14   Fatty   Acid-­‐Binding   Protein:   improvement   of   protein   stability   by   substitution   of   the   single   Cys62   residue,   Worldwide   Protein   Data   Bank.   Gill,   S.   C.,   and   Vonhippel,   P.   H.   (1989)   CALCULATION   OF   PROTEIN   EXTINCTION   COEFFICIENTS   FROM   AMINO-­‐ACID   SEQUENCE   DATA,   Analytical   Biochemistry   182,   319-­‐ 326.   Adams,  P.  D.,  Afonine,  P.  V.,  Bunkoczi,  G.,  Chen,  V.  B.,  Davis,  I.  W.,  Echols,  N.,  Headd,  J.   J.,   Hung,   L.   W.,   Kapral,   G.   J.,   Grosse-­‐Kunstleve,   R.   W.,   McCoy,   A.   J.,   Moriarty,   N.   W.,   Oeffner,  R.,  Read,  R.  J.,  Richardson,  D.  C.,  Richardson,  J.  S.,  Terwilliger,  T.  C.,  and  Zwart,   P.   H.   (2010)   PHENIX:   a   comprehensive   Python-­‐based   system   for   macromolecular   185     103.     104.     105.     106.     107.     108.     109.     110.     111.     112.   structure  solution,  Acta  Crystallographica  Section  D-­‐Biological  Crystallography  66,  213-­‐ 221.   Bailey,  S.  (1994)  THE  CCP4  SUITE  -­‐  PROGRAMS  FOR  PROTEIN  CRYSTALLOGRAPHY,  Acta   Crystallographica  Section  D-­‐Biological  Crystallography  50,  760-­‐763.   Adams,  P.  D.,  Afonine,  P.  V.,  Bunkoczi,  G.,  Chen,  V.  B.,  Echols,  N.,  Headd,  J.  J.,  Hung,  L.-­‐ W.,  Jain,  S.,  Kapral,  G.  J.,  Kunstleve,  R.  W.  G.,  McCoy,  A.  J.,  Moriarty,  N.  W.,  Oeffner,  R.   D.,   Read,   R.   J.,   Richardson,   D.   C.,   Richardson,   J.   S.,   Terwilliger,   T.   C.,   and   Zwart,   P.   H.   (2011)  The  Phenix  software  for  automated  determination  of  macromolecular  structures,   Methods  55,  94-­‐106.   Murshudov,  G.  N.,  Vagin,  A.  A.,  and  Dodson,  E.  J.  (1997)  Refinement  of  macromolecular   structures   by   the   maximum-­‐likelihood   method,   Acta   Crystallographica   Section   D-­‐ Biological  Crystallography  53,  240-­‐255.   Ong,   D.   E.,   and   Page,   D.   L.   (1987)   Cellular   retinol-­‐binding   protein   (type   two)   is   abundant   in  human  small  intestine,  Journal  of  Lipid  Research  28,  739-­‐745.   Haunerland,   N.   H.,   and   Spener,   F.   (2004)   Fatty   acid-­‐binding   proteins   –   insights   from   genetic  manipulations,  Progress  in  Lipid  Research  43,  328-­‐349.   Gillilan,   R.   E.,   Ayers,   S.   D.,   and   Noy,   N.   (2007)   Structural   Basis   for   Activation   of   Fatty   Acid-­‐binding  Protein  4,  Journal  of  Molecular  Biology  372,  1246-­‐1260.   Kaczocha,   M.,   Glaser,   S.   T.,   and   Deutsch,   D.   G.   (2009)   Identification   of   intracellular   carriers  for  the  endocannabinoid  anandamide,  Proceedings  of  the  National  Academy  of   Sciences  of  the  United  States  of  America  106,  6375-­‐6380.   Sulsky,  R.,  Magnin,  D.  R.,  Huanga,  Y.,  Simpkins,  L.,  Taunk,  P.,  Patel,  M.,  Zhu,  Y.,  Stouch,  T.   R.,  Bassolino-­‐Klimas,  D.,  Parker,  R.,  Harrity,  T.,  Stoffel,  R.,  Taylor,  D.  S.,  Lavoie,  T.  B.,  Kish,   K.,  Jacobson,  B.  L.,  Sheriff,  S.,  Adam,  L.  P.,  Ewing,  W.  R.,  and  Robl,  J.  A.  (2007)  Potent  and   selective   biphenyl   azole   inhibitors   of   adipocyte   fatty   acid   binding   pirotein   (aFABP),   Bioorganic  &  Medicinal  Chemistry  Letters  17,  3511-­‐3515.   Yapici,   I.,   Lee,   K.   S.   S.,   Berbasova,   T.,   Nosrati,   M.,   Jia,   X.   F.,   Vasileiou,   C.,   Wang,   W.   J.,   Santos,   E.   M.,   Geiger,   J.   H.,   and   Borhan,   B.   (2015)   "Turn-­‐On"   Protein   Fluorescence:   In   Situ   Formation   of   Cyanine   Dyes,   Journal   of   the   American   Chemical   Society   137,   1073-­‐ 1080.   Berbasova,  T.,  Nosrati,  M.,  Vasileiou,  C.,  Wang,  W.,  Lee,  K.  S.,  Yapici,  I.,  Geiger,  J.  H.,  and   Borhan,  B.  (2013)  Rational  design  of  a  colorimetric  pH  sensor  from  a  soluble  retinoic  acid   chaperone,  Journal  of  the  American  Chemical  Society  135,  16111-­‐16119.     186   113.     114.     115.     116.     117.     118.     119.     120.     121.     122.     Abraham,  M.  H.,  Grellier,  P.  L.,  Abboud,  J.  L.  M.,  Doherty,  R.  M.,  and  Taft,  R.  W.  (1988)   Solvent   Effects   in   Organic-­‐Chemistry   -­‐   Recent   Developments,   Canadian   Journal   of   Chemistry-­‐Revue  Canadienne  De  Chimie  66,  2673-­‐2686.   Ando,   Y.,   Homma,   Y.,   Hiruta,   Y.,   Citterio,   D.,   and   Suzuki,   K.   (2009)   Structural   characteristics   and   optical   properties   of   a   series   of   solvatochromic   fluorescent   dyes   displaying  long-­‐wavelength  emission,  Dyes  and  Pigments  83,  198-­‐206.   Carlotti,   B.,   Flamini,   R.,   Kikas,   I.,   Mazzucato,   U.,   and   Spalletti,   A.   (2012)   Intramolecular   charge   transfer,   solvatochromism   and   hyperpolarizability   of   compounds   bearing   ethenylene  or  ethynylene  bridges,  Chemical  Physics  407,  9-­‐19.   Kamlet,  M.  J.,  and  Taft,  R.  W.  (1976)  Solvatochromic  Comparison  Method  1.  Beta-­‐Scale   of  Solvent  Hydrogen-­‐Bond  Acceptor  (Hba)  Basicities,  Journal  of  the  American  Chemical   Society  98,  377-­‐383.   Yutani,   K.,   Ogasahara,   K.,   Tsujita,   T.,   and   Sugino,   Y.   (1987)   Dependence   of   Conformational   Stability   on   Hydrophobicity   of   the   Amino-­‐Acid   Residue   in   a   Series   of   Variant   Proteins   Substituted   at   a   Unique   Position   of   Tryptophan   Synthase   Alpha-­‐ Subunit,   Proceedings   of   the   National   Academy   of   Sciences   of   the   United   States   of   America  84,  4441-­‐4444.   Noguchi,  J.,  Chaen,  K.,  Nhuan  Thi,  V.,  Akasaka,  T.,  Shimada,  H.,  Nakashima,  T.,  Nishi,  A.,   Satoh,  H.,  Omori,  T.,  Kakuta,  Y.,  and  Kimura,  M.  (2011)  Crystal  structure  of  the  branching   enzyme  I  (BEI)  from  Oryza  sativa  L  with  implications  for  catalysis  and  substrate  binding,   Glycobiology  21,  1108-­‐1116.   Feng,   L.,   Fawaz,   R.,   Hovde,   S.,   Gilbert,   L.,   Chiou,   J.,   and   Geiger,   J.   H.   (2015)   Crystal   Structures   of   Escherichia   coli   Branching   Enzyme   in   Complex   with   Linear   Oligosaccharides,  Biochemistry  54,  6207-­‐6218.   Henrissat,  B.  (1991)  A  classification  of  glycosyl  hydrolases  based  on  amino  acid  sequence   similarities.,  Biochemical  Journal  280,  309-­‐316.   Jespersen,  H.  M.,  MacGregor,  E.  A.,  Henrissat,  B.,  Sierks,  M.  R.,  and  Sevensson,  B.  (1993)   Starch-­‐   and   Glycogen-­‐Debranching   and   Branching   Enzymes:   Prediction   of   Structural   Features   of   the   Catalytic   (Îą/β)8-­‐Barrel   Domain   and   Evolutionary   Relationship   to   Other   Amylolytic  Enzymes,  Journal  of  Protein  Chemistry  12,  791-­‐805.   Baba,   T.,   Kimura,   K.,   Mizuno,   K.,   Etoh,   H.,   Ishida,   Y.,   Shida,   O.,   and   Arai,   Y.   (1991)   SEQUENCE   CONSERVATION   OF   THE   CATALYTIC   REGIONS   OF   AMYLOLYTIC   ENZYMES   IN   MAIZE  BRANCHING  ENZYME-­‐I,  Biochem.  Biophys.  Res.  Commun.  181,  87-­‐94.   187   123.     124.     125.     126.     127.     128.     129.     130.     131.     132.     133.     134.   Boyer,   C.   D.,   and   Preiss,   J.   (1981)   Evidence   for   Independent   Genetic   Control   of   the   Multiple   Forms   of   Maize   Endosperm   Branching   Enzymes   and   Starch   Synthases,   Plant   Physicology  67,  1141-­‐1145.   Gao,   M.,   Fisher,   D.   K.,   Kim,   K.   N.,   Shannon,   J.   C.,   and   Guiltinan,   M.   J.   (1996)   Evolutionary   conservation  and  expression  patterns  of  maize  starch  branching  enzyme  I  and  llb  genes   suggests  isoform  specialization,  Plant  Molecular  Biology  30,  1223-­‐1232.   Guan,   H.,   Li,   P.,   Imparl-­‐Radosevich,   J.,   Preiss,   J.,   and   Keeling,   P.   (1997)   Comparing   the   Properties  of  Escherichia  coli  Branching  Enzyme  and  Maize  Branching  Enzyme,  Archives   of  biochemistry  and  biophysics  342,  92-­‐98.   Guan,   H.   P.,   Baba,   T.,   and   Preiss,   J.   (1994)   Expression   of   Branching   Enzyme   I   of   Maize   Endosperm  in  Escherichia  coli,  Plant  Physicology  104,  1449-­‐1453.   Guan,   H.   P.,   Kuriki,   T.,   Sivak,   M.,   and   Preiss,   J.   (1995)   Maize   Branching   Enzyme   Catalyzes   synthesis   of   Glycogen-­‐like   Polysaccharide   in   GlgB-­‐Deficient   Escherichia   Coli,   Proceedings   of  the  National  Academy  of  Sciences  in  the  United  States  of  America  92,  964-­‐967.   Guan,   H.   P.,   and   Preiss,   J.   (1993)   Differentiation   of   the   Properties   of   the   Branching   lsozymes  from  Maize  (Zea  mays),  Plant  Physiology  102,  1269-­‐1273.   Kuriki,   T.,   Guan,   H.,   Sivak,   M.,   and   Preiss,   J.   (1996)   Analysis   of   the   Active   Center   of   Branching  Enzyme  II  from  Maize  Endosperm,  J  Protein  Chem  15,  305-­‐313.   Kuriki,  T.,  Stewart,  D.  C.,  and  Preiss,  J.  (1997)  Construction  of  Chimeric  Enzymes  out  of   Maize   Endosperm   Branching   Enzymes   I   and   II::   ACTIVITY   AND   PROPERTIES,   Journal   of   Biological  Chemistry  272,  28999-­‐29004.   Takeda,   Y.,   Guan,   H.-­‐P.,   and   Preiss,   J.   (1993)   Branching   of   amylose   by   the   branching   isoenzymes  of  maize  endosperm,  Carbohhydr.  Res.  240,  253-­‐263.   Kawasaki,   T.,   Mizumo,   K.,   Baba,   T.,   and   Shimada,   H.   (1993)   Molecular   analysis   of   the   gene  encoding  a  rice  starch  branching  enzyme,  Molecular  and  General  Genetics  237,  10-­‐ 16.   Mizuno,  K.,  Kawasaki,  T.,  Shimada,  H.,  Satoh,  H.,  Kobayashi,  E.,  Okumura,  S.,  Arai,  Y.,  and   Baba,  T.  (1993)  Alteration  of  the  Structural  Properties  of  Starch  Components  by  the  Lack   of   an   Isoform   of   Starch   Branching   Enzyme   in   Rice   Seeds,   The   Journal   of   biological   chemistry  268,  19084-­‐19091.   Mizuno,   K.,   Kimura,   K.,   Arai,   Y.,   Kawasaki,   T.,   Shimada,   H.,   and   Baba,   T.   (1992)   Starch   Branching  Enzymes  from  Immature  Rice  Seeds,  Journal  of  Biochemistry  112,  643-­‐651.     188   135.     136.     137.     138.     139.     140.     141.     142.     143.     144.     Mizuno,   K.,   Kobayashi,   E.,   Tachibana,   M.,   Tsutomu,   K.,   Fujimura,   T.,   Funane,   K.,   Kobayashi,   M.,   and   Baba,   T.   (2001)   Characterization   of   an   isoform   of   rice   starch   branching  enzyme,  RBE4,  in  developing  seeds,  Plant  Cell  Physiol.  42,  349-­‐357.   Nakamura,  Y.,  Takeichi,  T.,  Kawaguchi,  K.,  and  Yamanouchi,  H.  (1992)  Purification  of  two   forms   of   starch   branching   enzyme   (Q-­‐enzyme)   from   developing   rice   endosperm,   Plant   Physiology  84,  329  –335.   Nakamura,  Y.,  Utsumi,  Y.,  Sawada,  T.,  Aihara,  S.,  Utsumi,  C.,  Yoshida,  M.,  and  Kitamura,   S.   (2010)   Characterization   of   the   reactions   of   starch   branching   enzymes   from   rice   endosperm,  Plant  and  Cell  Physiology  51,  776-­‐794.   Morell,   M.   K.,   Blennow,   A.,   Kosar-­‐Hashemi,   B.,   and   Samuel,   M.   (1997)   Differential   Expression   and   Properties   of   Starch   Branching   Enzyme   lsoforms   in   Developing   Wheat   Endosperm,  Plant  Physiol.  113,  201-­‐208.   Rahman,   S.,   Li,   Z.,   Abrahams,   S.,   Abbott,   D.,   Appels,   R.,   and   Morell,   M.   K.   (1999)   Characterisation   of   a   gene   encoding   wheat   endosperm   starch   branching   enzyme-­‐I,   Theor.  Appl.  Genet.  98,  156-­‐163.   Repellin,  A.,  BĂĽga,  M.,  and  Chibbar,  R.  N.  (2001)  Characterization  of  a  cDNA  encoding  a   type   I   starch   branching   enzyme   produced   in   developing   wheat   (Triticum   aestivum   L.)kernels,  Journal  of  Plant  Physiology  158,  91-­‐100.   Sadequr   Rahman*,   A.   R.,   Zhongyi   Li,   Yasuhiko   Mukai,   Maki   Yamamoto,   Behjat   Kosar-­‐ Hashemi,   Sharon   Abrahams,   and   Matthew   K.   Morell.   (2001)   Comparison   of   Starch-­‐ Branching   Enzyme   Genes   Reveals   Evolutionary   Relationships   Among   Isoforms.   Characterization  of  a  Gene  for  Starch-­‐Branching  Enzyme  IIa  from  the  Wheat  D  Genome   Donor  Aegilops  tauschii1,  Plant  Physiology  125.   Sun,  C.,  Sathish,  P.,  Ahlandsberg,  S.,  Deiber,  A.,  and  Jansson,  C.  (1997)  Identification  of   four   starch-­‐branching   enzymes   in   barley   endosperm-­‐   partial   purification   of   forms   I,   Ila   and  IIb,  New  Phytol.  137,  215-­‐222.   ChuanXin,   S.,   Sathish,   P.,   Deiber,   A.,   and   Jansson,   C.   (1995)   Multiple   starch   branching   enzymes   in   barley   endosperm,   Photosynthesis:   from   light   to   biosphere   Volume   V.   Proceedings  of  the  Xth  International  Photosynthesis  Congress,  Montpellier,  France,  317-­‐ 320.   Sun,   C.,   Sathish,   P.,   Ahlandsberg,   S.,   and   Jansson,   C.   (1998)   The   Two   Genes   Encoding   Starch-­‐Branching   Enzymes   IIa   and   IIb   Are   Differentially   Expressed   in   Barley,   Plant   Physiol.  118,  37–49.   189   145.     146.     147.     148.   149.     150.     151.     152.     153.     154.     155.     156.   Rydberg,   U.,   Andersson,   L.,   Andersson,   R.,   A  ̊man2,   P.,   and   Larsson,   H.   (2001)   Comparison  of  starch  branching  enzyme  I  and  II  from  potato,  Eur.  J.  Biochem.  268,  6140-­‐ 6145.   Visko-­‐Nielsen,   A.,   and   Blennow,   A.   (1998)   Isolation   of   starch   branching   enzyme   I   from   potato  using  gamma-­‐cyclodextrin  affinity  chromatography,  J.  Chromatogr.  800,  382-­‐385.   Salehuzzaman,   S.   N.   I.   M.,   Jacobsen,   E.,   and   Visser,   R.   G.   F.   (1992)   Cloning,   partial   sequencing   and   expression   of   a   cDNA   coding   for   branching   enzyme   in   cassava,   Plant   Molecular  Biology  20,  809-­‐819.   Amith,  A.  M.  (1988)  Major  differences  in  isoforms  of  starch-­‐branching  enzyme  between   developing  embryos  of  round-­‐  and  wrinkled-­‐seeded  peas  (Pisum  sativum  L.),  Planta  175,   270-­‐279.   Tomlinson,  K.  L.,  Lloydt,  J.  R.,  and  Smith,  A.  M.  (1997)  Importance  of  isoforms  of  starch-­‐ branching  enzyme  in  determining  the  structure  of  starch  in  pea  leaves,  The  Plant  Journal   11,  31-­‐43.   Zhu,  Z.  P.,  and  Chen,  S.  Q.  (1992,  20-­‐24  July)  Starch  accumulation  and  branching  enzyme   gene   expression   during   seed   development   of   soybean   and   pea,   Proceedings   of   the   Fourth   International   Workshop   on   Seeds:   basic   and   applied   aspects   of   seed   biology,   Angers,  France  1,  127-­‐133.   Nakamura,   Y.   (2002)   Towards   a   better   understanding   of   the   metabolic   system   for   amylopectin  biosynthesis  in  plants,  Plant  Cell  Physiol.  43,  718-­‐725.   Chaen,  K.,  Noguchi,  J.,  Omori,  T.,  Kakuta,  Y.,  and  Kimura,  M.  (2012)  Crystal  structure  of   the   rice   branching   enzyme   I   (BEI)   in   complex   with   maltopentaose,   Biochemical   and   Biophysical  Research  Communications  424,  508-­‐511.   Vu,  N.  T.,  Shimada,  H.,  Kakuta,  Y.,  Nakashima,  T.,  Ida,  H.,  Omori,  T.,  Nishi,  A.,  Satoh,  H.,   and  Kimura,  M.  (2008)  Biochemical  and  Crystallographic  Characterization  of  the  Starch   Branching  Enzyme  I  (BEI)  from  Oryz  sativa  L,  Biosci.  Biotechnol.  Biochem.  72,  2858-­‐2866.   Boyer,   C.   D.,   and   Preiss,   J.   (1978)   MULTIPLE   FORMS   OF   (1,4)-­‐a-­‐D-­‐GLUCAN,   (1,4)-­‐a-­‐D-­‐ GLUCAN-­‐6-­‐   GLYCOSYL   TRANSFERASE   FROM   DEVELOPING   Zea   mays   L.   KERNELS,   Carbohydrate  Research  61,  321-­‐334.     Hizukuri,  S.,  Takeda,  Y.,  Yasuda,  M.,  and  Suzuki,  A.  (1981)  MULTI-­‐BRANCHED  NATURE  OF   AMYLOSE   AND   THE   ACTION   OF   DE-­‐   BRANCHING   ENZYMES,   Carbohydrate   Research   94,   205-­‐213.   Binderup,   K.,   and   Preiss,   J.   (1998)   Glutamate-­‐459   Is   Important   for   Escherichia   coli   Branching  Enzyme  Activity,  Biochemistry  37,  9033-­‐9037.   190   154.     155.     156.     156.       Boyer,   C.   D.,   and   Preiss,   J.   (1978)   MULTIPLE   FORMS   OF   (1,4)-­‐a-­‐D-­‐GLUCAN,   (1,4)-­‐a-­‐D-­‐ GLUCAN-­‐6-­‐   GLYCOSYL   TRANSFERASE   FROM   DEVELOPING   Zea   mays   L.   KERNELS,   Carbohydrate  Research  61,  321-­‐334.     Hizukuri,  S.,  Takeda,  Y.,  Yasuda,  M.,  and  Suzuki,  A.  (1981)  MULTI-­‐BRANCHED  NATURE  OF   AMYLOSE   AND   THE   ACTION   OF   DE-­‐   BRANCHING   ENZYMES,   Carbohydrate   Research   94,   205-­‐213.   Binderup,   K.,   and   Preiss,   J.   (1998)   Glutamate-­‐459   Is   Important   for   Escherichia   coli   Branching  Enzyme  Activity,  Biochemistry  37,  9033-­‐9037.   Binderup,   K.,   and   Preiss,   J.   (1998)   Glutamate-­‐459   Is   Important   for   Escherichia   coli   Branching  Enzyme  Activity,  Biochemistry  37,  9033-­‐9037.     191 Â